long-term trends in aquatic pollutants: chloride and ... · long-term trends in aquatic pollutants:...

187
Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds by Amy Marie Kamarainen A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Zoology) at the University of Wisconsin, Madison 2009

Upload: others

Post on 30-Jun-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in

lakes embedded in urban and agricultural watersheds

by

Amy Marie Kamarainen

A dissertation submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

(Zoology)

at the

University of Wisconsin, Madison

2009

Page 2: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

UMI Number: 3384162

INFORMATION TO USERS

The quality of this reproduction is dependent upon the quality of the copy

submitted. Broken or indistinct print, colored or poor quality illustrations

and photographs, print bleed-through, substandard margins, and improper

alignment can adversely affect reproduction.

In the unlikely event that the author did not send a complete manuscript

and there are missing pages, these will be noted. Also, if unauthorized

copyright material had to be removed, a note will indicate the deletion.

UMI UMI Microform 3384162

Copyright 2009 by ProQuest LLC All rights reserved. This microform edition is protected against

unauthorized copying under Title 17, United States Code.

ProQuest LLC 789 East Eisenhower Parkway

P.O. Box 1346 Ann Arbor, Ml 48106-1346

Page 3: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

3 A dissertation entitled

•a e

Long-term trends in aquatic pollutants: Chloride

O

V BO cd D.

and phosphorus dynamics of lakes embedded in

urban and agricultural watersheds

o U

submitted to the Graduate School of the University of Wisconsin-Madison

in partial fulfillment of the requirements for the degree of Doctor of Philosophy

by

Amy Marie Kamarainen

Date of Final Oral Examination: 11 May 2009

Month & Year Degree to be awarded: December May August 2 0 f J 9

O

O c

A * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

Approval Signatures of Dissertation Committee

S-fay W~ (s Gya?Azgf~

^im/M A'(TtuMJU^

Signature, Dean of Graduate School

K^n CkjlmJhJtx- 4tJ

Page 4: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

TABLE OF CONTENTS ii

Page

Abstract iii

Acknowledgements v

Introduction 1

Chapter 1: Road salt management and chloride dynamics 12

of an urban lake

Chapter 2: Phosphorus sources and demand during summer in a 52

eutrophic lake

Chapter 3: Estimates of phosphorus entrainment in Lake Mendota: 98

A comparison of one-dimensional and three-dimensional approaches

Chapter 4: Long-term trends in ice cover, stability, phosphorus and 141

water quality in eutrophic Lake Mendota

Page 5: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

iii ABSTRACT

Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in

lakes embedded in urban and agricultural watersheds

Amy Marie Kamarainen

Under the supervision of Professor Stephen R. Carpenter

At the University of Wisconsin, Madison

Aquatic ecosystems are increasingly affected by human activities such as land

use change in the watershed. Changes in land use affect the delivery of non-point

source loads of nutrients and pollutants. I explore long-term trends and drivers of two

aquatic pollutants, chloride and phosphorus, associated with changes in land use and

land cover. Comparison of these two pollutants provided an opportunity to explore

simple and complex models for solute transport and processing in lakes embedded in

changing landscapes.

Excessive chloride loading has become pervasive due to increases in road miles

and rates of road salt application associated with urbanizing watersheds in northern

climates. In the first chapter, I examined the relationship between road salt application

and chloride dynamics in an urban watershed, Lake Wingra. Using a model calibrated

with long-term data, I explored scenarios for changes in road salt management. I found

that, under current conditions, the lake should quickly respond to changes in road salt

application and projected mean concentrations are unlikely to exceed guidelines for

Page 6: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

iv aquatic organisms. However, trends in chloride concentrations of the groundwater

underlying the Wingra watershed suggested a need to mitigate road salt application.

In the second portion of my dissertation, I explored phosphorus dynamics in Lake

Mendota to assess the relative importance of external loading and recycling of this

limiting nutrient. Using a combination of field sampling, long-term data analysis and

biological and physical modeling, I assessed the contribution of three different

mechanisms of phosphorus recycling: biotic recycling, entrainment and sediment

release. Both biotic recycling and entrainment provided important sources of P to

primary producers during the stratified season. Entrainment was spatially and

temporally variable, but could be reasonably represented using both a 3-D

hydrodynamic model and sampling based on a single central location. Finally, although

long-term P dynamics were primarily driven by external loading, P in the hypolimnion

increased over time. This change was driven by changes in the stability of the water

column and increased length of stratification. Changes in the physical condition of the

lake held implications for phosphorus recycling and water clarity during the summer.

Page 7: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

V

ACKNOWLEDGEMENTS

I would like to begin by thanking my adviser, Steve Carpenter, for all he has done

to mold me into the budding scientist and educator that I have become. With even-

handed guidance, encouragement, and intellectual latitude he has turned advising into a

well-honed craft. Perhaps most transferable among the skills he nurtured are those of

time management and an eye for what is at the heart of any scientific pursuit. His ability

to sift and winnow is unmatched, and I can only hope that I've picked up a fraction of his

skill in this area. I am extremely grateful for his time and guidance; for the lab trips to the

Arb and cabin; for discussions of Tufte; and for bringing together so many keen and

thoughtful people in the lab.

The members of my committee deserve heaps of thanks for their support of and

interest in my projects and professional development. Dave Armstrong was a great

instructor and shared insight from many years of experience studying the chemistry of

Madison lakes - it was always great to run into Dave at the Union on a coffee run. Emily

Stanley shared friendly advice on topics ranging from environmental chemistry to grad

school, to life; I sought and valued her opinion on all of the above. I always tried to be

more of a landscape ecologist than my thesis would allow - and much of my

enthusiasm stems from Monica Turner's broad and engaging approach to the topic.

Chin Wu had an amazing capacity for collaboration and I enjoyed working with him and

his students throughout my degree.

It was a pleasure to interact with the members of the Long-Term Ecological

Research (LTER) Project throughout my tenure. The breath of people and fields

Page 8: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

vi

represented in this group was refreshing and inspiring. I am also indebted to the LTER

program for generous financial support to attend meetings and conduct research during

the summers. Another source of funding I feel honored to have received was from the

National Science Foundation Graduate Research Fellowship program. Finally, the

Zoology Department consistently supported me as I traveled to meetings and

conferences. I would not have been able to make so many connections and have so

many rich learning experiences without these generous supporters.

I can't say enough about the friends and colleagues I've met at the CFL. There

was my cohort: Olaf Jensen, Stephanie Schmidt, Nico Preston, Matt Van de Bogert, Jeff

Waters, David Gilroy - it was great to be thrown into the crazy world that is grad-student

life with such fun-loving, hard-working and inspiring people (there will be a reunion at

Wando's in 2014). Steph deserves a special shout ou t - I don't think I would have made

it through the rough patches without your friendship and support. There were many CFL

grads who welcomed me with open arms: Cailin Gille, Chris Solomon, Matt Diebel,

Brian Roth, Greg Sass, Norman Mercado-Silva, Piet Johnson, Caitlin Orr, Abby Popp,

Elena Bennet, Kristy Rogers - thanks for showing me the ropes, helping me learn how

to say no, and reminding me that sometimes you just need to leave limnology at the lab

(but apparently brain-shaped jello molds are still ok). And to those who came next: Brian

Weidel, Katrina Butkas, Noah Lottig, Justin Fox, Oonsie Biggs, Matt Fuller, Matt Kornis,

Jennifer Schmitz, Owen Langman, Daniel Collins, Julian Olden, Gretchen Hansen,

Jereme Gaeta, Steve Powers, Marit Sallstrom, Mona Papes, Ishi Buffam, Scott Higgins,

Erika Nilsson, and Luke Winslow - thank you all for your friendship and collegiality. I

Page 9: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

vii

looked forward to LAFS and Science Club nights and benefitted from the open and

honest conversations that would come out of these gatherings. It was great to have

folks who participated from all across campus - thanks for coming down Stuart Jones,

Ashley Shade, Emily Kara and Eric Booth!

My time in Madison was enriched by folks in many organizations across campus.

Thanks to Peggy Nowicki, Joan Ersland, Sharon Kahn, Will Egen and Gale Oakes in the

Zoo Department - there were many friendly faces during my visits to Noland Hall. Also,

thanks to all the folks in GSIS for great lunch-time topics of conversation and for keeping

me connected (Grace Lee, Genny Kozak, Sarah Heimovics, Sarah Alger!). To Tessa

Desmond-Lowinski, Brian Manske and Don Gillian-Daniels at Delta - thanks for

supporting me throughout the certificate experience. I feel lucky to have been able to

rely on your support and guidance -1 couldn't have done it without your kindness and

positive attitudes. Folks from the Friends of Lake Wingra provided support, old data,

great anecdotes and gave me a great source of inspiration and motivation - thank you

Jim Lorman, Anne Forbed, Kirsti Sorsa and David Leibl.

To the folks that kept me company during the summers at the Mad Lakes Lab -

Ted Bier, James Thoyre and Mark Lochner I couldn't have gotten through my first field

season (much less the rest of grad school) without your help. Thanks to Carol

Schraufnagel, Denise Karns, Trish Haza, Anne Murphy-Lomm, Mary Possin, and

Marilyn Larsen who helped me slash through the jungles of the administrative world -

and they made it look easy. Thanks to Dave Balsiger and Barbara Benson who helped

me get organized and taught me many tricks for accessing and analyzing the incredible

Page 10: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

viii

LTER dataset. A huge thank you to Paul Hanson for many helpful conversations and for

working with me to decipher buoy data in those early years rife with tipping and

technical difficulties. Jeff Maxted was an indispensable source for anything GIS - I was

amazed by and grateful for his level of skill and his willingness to share your knowledge

with us all. "Thank you" doesn't seem enough for Dick Lathrop - he helped me wrap my

head around the Madison Lake and helped me see things through the lens of lake

management. I am very appreciative for all the insight and long-term data that he

shared and showed me how to harness.

I have a dept of gratitude that cannot possibly be paid to the undergraduate

students who helped in the field and in the lab. It was a pleasure to work with Leif

Evensen on the chloride project. Freya Rowland gave her all and then some during the

Zooplankton Biomass field campaign of 2006. Her enthusiasm in the field was

contagious and her persistence on the scope was admirable. Rachel Penczykowski may

be the only person who could match Freya's enthusiasm in the field. She was my

sidekick for three years - and I was sad to let her go when she graduated. We worked

hard and had fun developing new lyrics for the latest limnology pop song or show tune.

There is no one I would trust more to run my SRPs!

I also am eternally grateful for friends and family who put up with my busy

schedule and helped me remember what is important. My parents, Pat and Jean, have

been my strongest supporters throughout the process. They were behind me all the way

- during my undergrad when I seemed to be building my resume based on the most

bizarre creature I could find to study during the summer and when I came to grad school

Page 11: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

ix

instead of finding a "real job". I am so grateful for their love and support and their

frequent offers to help out where they could. Weekly conversations with my sister Jen

and brother Steve helped me stay grounded - and family vacations were the highlight of

each winter. I appreciate all the times they came to visit me when I felt too busy to

travel. Chiana and Megan are source for laughs and shoulders should I ever need them.

It was and is so comforting to have such good friends.

I also want to extend a special thank you to Nicholas Preston. Thanks for coffee

at Panera, great conversation wherever we go, getting me into ski patrol and for

opening my eyes and my world in so many ways. The grad school experience was

made so much better with you by my side. Your love and encouragement have meant

the world to me.

Page 12: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

INTRODUCTION 1

Aquatic ecosystems are increasingly affected by human activities. Among the

most significant pathway of influence is through runoff of nutrients and pollutants from

the surrounding landscape. The influence of watershed composition on lake

characteristics has long been recognized (Dillon and Kirchner 1974, O'Sullivan

1979). Extreme changes in land use and land cover result in changes in hydrology

and solute export from the landscape (Bormann and Likens 1967, Likens et al. 1970).

Nonpoint pollution - runoff of nutrients and toxins through diffuse landscape

pathways rather than isolated discharge points - is now the principal driver of

freshwater eutrophication in developed countries (Carpenter et al. 1998). Extensive

changes in land use and land cover are occurring and will continue to occur globally

in order to meet the food and housing needs of a growing global population (Meyer

and Turner 1994, Millenium Ecosystem Assessment 2005, Foley et al. 2005).

Agriculture and urban land use have become the principal endpoints of land

use change. According to a United Nations report, 2008 was the first year in which

over 50% of the world's population lived in urban areas, and this trend is projected to

continue (UN-HABITAT 2008). Conversion of native land to residential and

commercial use is associated with an increase in roads and impervious surface in the

watershed (Arnold and Gibbons 1996). Additionally, urban roadways are generally

underlain by extensive stormwater drainage networks that efficiently convey runoff to

nearby retention basins or natural aquatic ecosystems (Smith et al. 1998, Hatt et al.

Page 13: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

2 2004). The net effect of urbanization is a dramatically altered hydrologic regime that

may readily transport pollutants to urban lakes and streams (Paul and Meyer 2001,

Brabec et al. 2002).

The impact of urbanization on aquatic ecosystems is surpassed by that of

agriculture, which is the single most significant cause of impaired waters in the United

States (USEPA 2002). The negative influence of agriculture on aquatic resources is

primarily driven by the redistribution of nutrients, due to use of synthetic fertilizers and

manure, and subsequent runoff of nutrient-rich waters and sediment into lakes and

streams (Bennett et al. 1999, Tilman et al. 2001). Over the past 40 years there has

been a 12% increase in global cropland, and a 700% increase in fertilizer use (Foley

et al. 2005, Tilman et al. 2001). The number of acres of land in agricultural production

is projected to increase in developing nations by at least 23%, but may decrease in

the developed world (Balmford 2005).

This dissertation is motivated by a need to better understand processes

driving long-term trends in aquatic pollutants associated with changing landscapes. I

focused on two very different solutes, chloride and phosphorus, which are important

nonpoint pollutants in the Madison area. I used a combination of field sampling,

modeling and long-term data analysis to assess the rates and dominant pathways of

in-lake processing of phosphorus coming from the predominantly agricultural

watershed of Lake Mendota. I also used long-term data analysis and modeling to

explore scenarios for road salt management and chloride dynamics in the highly

urbanized watershed of Lake Wingra.

Page 14: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

In the 1950s, many municipalities began applying rock salt to roads during

the winter to improve driving conditions. The long-term effects of road salt application

are now being recognized due to an increase in chloride and sodium concentrations

in lakes, streams and groundwaters across northern latitudes (Kaushal et al. 2002,

Jackson and Jobbagy 2005, Howard and Maier 2007, Novotny et al. 2008). A number

of studies document increasing chloride concentrations over time, yet few offer

assessment of how chloride levels may respond to mitigation efforts (Novotny et al.

2008). In Chapter 1,1 use a long-term dataset from the Lake Wingra watershed to

characterize the drivers of chloride dynamics, build a model to represent changes in

chloride concentration, and assess how the system may respond to changes in road

salt management. Chloride is a conservative tracer, meaning that the mass added to

a water body stays constant and is not altered by reactions with other chemical

species or biotic uptake or release. Because of its conservative nature, I was able to

represent the long-term changes in chloride concentrations using a simple model.

Using this simple model I show that Lake Wingra should respond quickly to changes

in chloride load and that current loading rates are not likely to exceed toxicity limits for

aquatic organisms. However, the capacity of the lake to respond quickly to changes

in load may be dramatically altered by accumulation of chloride in shallow

groundwaters. Accumulation of chloride in groundwaters raises a number of concerns

for aquatic resources in the Madison area, which I explore fully in Chapter 1.

Whereas chloride is a conservative pollutant, the second pollutant of interest,

phosphorus, is biologically and chemically reactive, and its concentrations are

Page 15: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

4 affected by uptake and release within aquatic ecosystems. Phosphorus (P) is

frequently the macronutrient that limits primary production in freshwater ecosystems,

so that biologically available P (PO43) is quickly taken up by plankton (Schindler

1977). Additionally, P readily forms complexes with iron, aluminum or calcium in the

sediments of a lake (Golterman 1973, Baccini 1985). These complexes may render

the P biologically inactive, effectively trapping P in lake sediments, until redox

conditions become favorable for the chemical release of P into the water column

(Mortimer 1941). The term "P recycling" has been used to describe processes by

which bound P is made available for biological uptake. Mechanisms of recycling may

include: 1.) recycling among biotic components through a cycle of consumption,

excretion and uptake (Vanni 2002), 2.) release of phosphorus within the hypolimnion

(from the sediments or sedimenting material) due to chemical or bacterial processes

(Lee et al. 1977, Marsden 1989), or 3.) mixing of phosphorus-rich hypolimnetic water

into the epilimnion, which is also called entrainment (Soranno et al. 1997, Stauffer

1987) (Figure 1). Of course, a build-up of P in the hypolimion via mechanism #2 is a

prerequisite for mechanism #3. Recycling mechanisms may contribute significantly to

the annual P budget and may be responsible for delayed lake recovery following

mitigation of external loading (Soranno et al. 1997, Nurnberg and Peters 1984,

Jeppesen et al. 1998, Sondergaard et al. 2001, Jeppesen et al. 2005). Given the

complex chemical and biological interactions that are relevant to phosphorus cycling

in lakes, I needed to use a more complex, multi-faceted approach to explore the

Page 16: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

5 importance of P recycling relative to external loading in Lake Mendota (Chapters 2,

3 and 4).

Lake Mendota is situated in an agricultural watershed where excessive

phosphorus loading has led to a perpetual eutrophic state, yet during the summer,

biologically available phosphorus falls below detection limits and is assumed to be

limiting primary production in the epilimnion. Despite extremely low concentrations of

bio-available P, phytoplankton communities maintain positive rates of primary

production (Kamarainen et al. 2009a). In Chapter 2, I explored whether external

load, entrainment from the hypolimnion, or biotic recycling within the epilimnion was

most likely providing the P necessary for the observed rates of primary production.

To do so, I combined a model of whole ecosystem metabolism with empirical

measurements of the major components of the phosphorus budget during the

summer of 2007. Findings suggest that biotic recycling and entrainment both

contribute P necessary to sustain primary production (Kamarainen et al. 2009a).

Having confirmed the importance of entrainment in biological production, I

combined an intensive field sampling campaign with a calibrated 3-D hydrodynamic

model of Lake Mendota to assess the spatial and temporal variability in phosphorus

entrainment (Chapter 3). Using this approach, I found that entrainment was spatially

variable. Nonetheless, sampling at a single location could estimate the mean

magnitude of the process (Kamarainen et al. 2009b). Also, the 3-D model

successfully represented the spatial variation in phosphorus concentrations, thus

Page 17: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

6 supporting the use of this model as a tool to accurately represent the physical

dynamics relevant in the transport of solutes within the basin.

In Chapter 4, I examined long-term changes in ice cover, physical

characteristics, phosphorus dynamics and water quality metrics in Lake Mendota

over thirty years. There was a significant decline in the ice cover and increase in the

strength and duration of stratification. At the same time phosphorus concentrations

increased and dissolved oxygen concentrations decreased in the hypolimnion of the

lake. The temporal synchrony between P and oxygen conditions in the hypolimnion

was likely driven by a common relationship with increasing stability and length of

stratification over time. Changes in the physical and chemical features of the lake

were tied to changes in water quality. We observed a significant improvement in

water clarity over time and a direct relationship between mean phosphorus

concentrations in the epilimnion and stability of the water column. While hypolimnetic

P concentrations increased, metrics of internal loading were, at best, weak predictors

of residual variation in April P concentrations predicted by the Vollenweider model.

Therefore, long-term P dynamics in the system were driven primarily by external

loads. In conclusion, water quality and phosphorus dynamics in Lake Mendota have

been affected by changes in ice cover and the strength and duration of stratification.

Yet, reductions of external loading, not further climate change, are the key to

restoration of water quality.

In summary, this thesis combines long-term data, short-term field campaigns,

and ecosystem models to analyze the dynamics of a conservative tracer, chloride,

Page 18: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

7 and a highly reactive substance, phosphorus, in lakes. A relatively simple model

successfully explained chloride dynamics. However, ongoing chloride enrichment of

groundwater, which was not addressed by my model, could complicate chloride

dynamics while greatly increasing year-round chloride concentrations in Lake Wingra

in the future. Phosphorus dynamics, in contrast, were more complicated because of

high rates of uptake and release within Lake Mendota. To understand phosphorus

dynamics, I used statistical models of a 30-year time series in combination with sub-

annual models of phosphorus uptake and recycling, including a detailed 3-D

hydrodynamic model. Taken together, my analyses showed that recycling of

phosphorus in Lake Mendota is biologically important during the stratified season and

is related to recent P inputs (on a time frame of 1-2 years) as well as physical stability

of the water column.

Page 19: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

References 8

Arnold, C. L, and C. J. Gibbons. 1996. Impervious surface coverage - The emergence of a key environmental indicator. Journal of the American Planning Association 62: 243-258.

Baccini, P. 1985. Phosphate interactions at the sediment-water interface. In: Stumm, W. [ed.], Chemical processes in lakes. Wiley.

Bennett, E. M., T. Reed-Andersen, J. N. Houser, J. R. Gabriel, and S. R. Carpenter. 1999. A phosphorus budget for the Lake Mendota watershed. Ecosystems 2: 69-75.

Bormann, F. H., and G. E. Likens. 1967. Nutrient cycling. Science 155: 424-&.

Brabec, E., S. Schulte, and P. L. Richards. 2002. Impervious surfaces and water quality: A review of current literature and its implications for watershed planning. Journal of Planning Literature 16: 499-514.

Carpenter, S. R., N. F. Caraco, D. L. Correll, R. W. Howarth, A. N. Sharpley, and V. H. Smith. 1998. Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecol Appl 8: 559-568.

Dillon, P. J., and W. B. Kirchner. 1975. Effects of geology and land-use on export of phosphorus from watersheds. Water Research 9:135-148.

Foley, J. A. and others 2005. Global consequences of land use. Science 309: 570-574.

Golterman, H. L. 1973. Vertical movement of phosphate in freshwater. In: Griffith, E. J. [ed.], Environmental phosphorus handbook. Wiley.

Hatt, B. E., T. D. Fletcher, C. J. Walsh, and S. L Taylor. 2004. The influence of urban density and drainage infrastructure on the concentrations and loads of pollutants in small streams. Environ Manage 34:112-124.

Howard, K. W. F., and H. Maier. 2007. Road de-icing salts as a potential constraint on urban growth in the Greater Toronto Area, Canada. Journal of Contaminant Hydrology 91:146-170.

Jackson, R. B., and E. G. Jobbagy. 2005. From icy roads to salty streams. P Natl Acad Sci USA 102: 14487-14488.

Page 20: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

9 Jeppesen, E., M. Sondergaard, J. P. Jensen, E. Mortensen, A. M. Hansen, and T. Jorgensen. 1998. Cascading trophic interactions from fish to bacteria and nutrients after reduced sewage loading: An 18-year study of a shallow hypertrophic lake. Ecosystems 1:250-267.

Jeppesen, E. and others 2005. Lake responses to reduced nutrient loading - an analysis of contemporary long-term data from 35 case studies. Freshwater Biol 50: 1747-1771.

Kamarainen, A. M., R. M. Penczykowski, M. C. Van de Bogert, P. C. Hanson, S. R. Carpenter. 2009a. Phosphorus sources and demand during summer in a eutrophic lake. Aquatic Sciences 71:214-227.

Kamarainen, A. M., H. Yuan, C. Wu, S. R. Carpenter. 2009b. Estimates of phosphorus entrainment in Lake Mendota: A comparison of one-dimensional and three-dimensional approaches. Limnology and Oceanography Methods 7:553-567.

Lee, G. F., W. C. Sonzogni, R. D. Spear. 1977. Significance of oxic vs anoxic conditions for Lake Mendota sediment phosphorus release. In: Golterman, H. L. [ed.], Interactions between sediments and fresh water: proceedings of an international symposium held at Amsterdam, the Netherlands, September 6-10, 1976. W. Junk.

Likens, G. E., F. H. Bormann, N. M. Johnson, D. W. Fisher, and R. S. Pierce. 1970. Effects of forest cutting and herbicide treatment on nutrient budgets in Hubbard Brook Watershed-Ecosystem. Ecol Monogr 40: 23-.

Marsden, M. W. 1989. Lake restoration by reducing external phosphorus loading -the influence of sediment phosphorus release. Freshwater Biol 21:139-162.

Meyer, W. B., B. L Turner, and University Corporation of America. 1994. Changes in land use and land cover: a global perspecative : papers arising from the 1991 OIES Global Change Institute. Cambridge University Press.

Millenium Ecosystem Assessment. 2005. Ecosystems and human well-being: current state and trends. Island Press, Washington, D.C., USA.

Mortimer, C. H. 1941. The exchange of dissolved substances between mud and water in lakes. Journal of Ecology 29: 280-329.

Novotny, E. V., D. Murphy, and H. G. Stefan. 2008. Increase of urban lake salinity by road deicing salt. Sci Total Environ 406:131-144.

Page 21: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

10 Nurnberg, G., and R. H. Peters. 1984. Biological availability of soluble reactive phosphorus in anoxic and oxic fresh-waters. Canadian Journal of Fisheries and Aquatic Sciences 41: 757-765.

O'Sullivan, P. E. 1979. The ecosystem-watershed concept in the environmental sciences - a review. International Journal of Environmental Studies 13: 273-281.

Paul, M. J. and J. L. Meyer. 2001. Streams in the urban landscape. Annual Review of Ecology and Systematics 32:333-365.

Schindler, D. W. 1977. Evolution of phosphorus limitation in lakes. Science 195: 260-262.

Smith, D. W., R. M. Facey, V. Novotny, and D. A. Kuemmel. 1998. Management of winter diffuse pollution from urban areas: Effect of drainage and deicing operations, p. 243-257. In: D. E. Newcomb [ed.], Ninth International Conference on Cold Regions Engineering. American Society of Civil Engineers.

Sondergaard, M., J. P. Jensen, and E. Jeppesen. 2001. Retention and internal loading of phosphorus in shallow, eutrophic lakes. The Scientific World 1: 427-442.

Soranno, P. A., S. R. Carpenter, and R. C. Lathrop. 1997. Internal phosphorus loading in Lake Mendota: response to external loads and weather. Canadian Journal of Fisheries and Aquatic Sciences 54:1883-1893.

Stauffer, R. E. 1987. Effects of oxygen transport on the areal hypolimnetic oxygen deficit. Water Resour Res 23:1887-1892.

Tilman, D. and others 2001. Forecasting agriculturally driven global environmental change. Science 292: 281-284.

United Nations Human Settlement Programme (UN-HABITAT). 2008. State of the worlds cities: Harmonious cities. Earthscan, London, UK.

United States Environmental Protection Agency (USEPA). 2002. National water quality inventory: Report to Congress. EPA 841 -R-07-001

Vanni, M. J. 2002. Nutrient cycling by animals in freshwater ecosystems. Annu Rev EcolSyst33:341-370.

Page 22: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

11

1.) B10TIC

A

2.) RELEASE FROM SEDIMENTS (Internal Load)

3.) ENTRAINMENT

MECHANISMS OF PHOSPHORUS

RECYCLING

Figure 1. Conceptual diagram of the potential mechanisms of phosphorus recycling within lakes. Biotic recycling (1) refers to cycling of phosphorus among biotic components within the epilimnion. Release from sediments (2) represents the release, by chemical or bacterial means, of phosphorus derived from the sediments or freshly sedimented organic material. Entrainment (3) refers to mixing of phosphorus-rich hypolimnetic water across the thermocline into the epilimnion.

Page 23: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

12

CHAPTER 1

Road salt management and chloride dynamics of an urban lake

by

Amy M. Kamarainen1,4 , Stephen R. Carpenter1, James Lorman2, and Richard C.

Lathrop13

1 Center for Limnology, University of Wisconsin, Madison, 680 N. Park St., Madison,

Wl53706

2 Department of Natural Science, Edgewood College, 1000 Edgewood College Drive,

Madison, Wl 53711

3 Wisconsin Department of Natural Resources, 1350 Fermite Drive, Monona, Wl

53716

4 Corresponding author: [email protected]

Keywords: chloride, road salt, urban watershed, lake, environmental management

Status: In preparation for the Journal of Environmental Quality

Page 24: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

13

Abstract

Chloride concentrations have been increasing in freshwaters across the United

States and Canada. These increases are widely attributed to the use of rock salt as a

de-icing agent. In a given watershed, changes in rock salt application could be

related to changes in urbanization, length and width of roads, weather, or road

management practices. The roles of these factors must be assessed in order to

predict or manage future changes in chloride concentrations. Here we develop a

model based on road salt application rates, weather and hydrology to explain long-

term (1973 - 2006) chloride dynamics in a shallow urban lake, Lake Wingra,

Wisconsin. We use this model to explore consequences of different road salt

application rates for chloride concentrations in the lake. A relatively simple model

satisfactorily represented monthly chloride dynamics in Lake Wingra, but tended

toward conservative estimates of extreme values. Simulations demonstrated that

Lake Wingra responds rapidly to changes in chloride loading. Reductions in road salt

use can mitigate chloride concentrations in the lake within ~5 years. Mean steady-

state chloride concentrations are not projected to exceed the limits for health of

humans and aquatic organisms, however, seasonal peak concentrations in excess of

the United States Environmental Protection Agency chloride guidelines already occur.

Groundwater contamination could increase chloride loadings to the lake and

decrease the lake's capacity to respond to reductions in road salt use. Uncertainty,

especially in the rate of increase and time for recovery of chloride concentrations in

Page 25: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

14 groundwater, highlight the need for detailed monitoring of road salt application and

pathways of interaction between surface and groundwater resources.

Page 26: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

15 Introduction

Road salt use is positively correlated with chloride concentrations in surface

and groundwater across Canada and northern parts of the United States (Schindler

2000, Interlandi and Crockett 2003, Kaushal et al. 2005, Howard and Maier 2007).

Increasing chloride concentrations have been most prevalent in urban and suburban

watersheds with a high incidence of impervious surface (Siver et al. 1999, Novotny et

al. 2008, Kelly 2008). Kaushal and others (2005) identified an exponential

relationship between mean chloride concentrations and the percent of impervious

surface in the watershed. Elevated chloride concentrations are not confined to urban

areas, but have also been identified in rural areas where road salt is applied to major

roadways during the winter (Schindler 2000). It is clear that the application of rock

salt to roadways in northern climates is affecting chloride concentrations of water

resources.

While the relationship between urbanization (as quantified by the number of

road miles or percent of impervious surface) and increasing chloride concentrations

is clear, there remains substantial variation in observed chloride concentrations that

cannot be explained by changes in land use alone. Variation among lakes has been

attributed to the watershed area:lake volume ratio, which serves as a proxy for the

flushing rate of the lake (Novotny et al. 2008). Also, land-cover permeability and

mode of storm water management within the watershed affect delivery of chloride to

a lake. For example, in the Toronto area, up to 55% of road salt applied may enter

the shallow groundwater system rather than being directly transported to surface

Page 27: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

16 waters. Meanwhile, Smith and others (1998) show that drainage by storm sewers

(as opposed to combined sewers or roadside swales) provides the most direct route

for pollutant transfer to surface waters. Thus, the hydrogeology of the watershed is

an important determinant of the fate and transport of chloride to surface and

groundwater (Lindstrom 2005).

Detailed accounts of chloride dynamics within watersheds show seasonal

peaks in chloride concentrations associated with snowmelt and spring rains (Koryak

et al. 2001, Novotny et al. 2008). Also, inter-annual variation in weather conditions

affects both road salt application rates and the water balance (Lindstrom 2005,

Thunqvist 2003). If we hope to improve management of road salt and curb the rise in

chloride concentrations, we need to better understand the drivers of seasonal and

inter-annual variation of chloride concentrations within and among lakes, streams and

aquifers.

Here we aim to improve understanding of drivers of inter-annual variation in

chloride concentration using twenty-four years of monitoring data from an urban

watershed. In this watershed there has been little change in road density since the

early 1970s, yet road salt application and chloride concentrations in the lake have

been steadily increasing. Thus, the effects of meteorological factors, road salt

management practices, and hydrological variables may be viewed independently

from changes in road density in the watershed. This provides an opportunity to study

a system that represents a possible future for many currently urbanizing watersheds.

We first assessed which combination of meteorological, hydrological and road salt

Page 28: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

17 application variables could best describe observed chloride dynamics in the lake.

Then, using the best-fitting model, we explored how changes in road salt

management are likely to affect chloride dynamics and we quantified the rate of lake

response to these changes.

Materials and methods

Lake Wingra is a small lake (area = 1.3 km2, mean depth = 2.7 m) situated in

an urban watershed (area = 14 km2) underlain by an extensive storm water sewer

network. The watershed is almost exclusively under the jurisdiction of the City of

Madison, Wisconsin, USA (43.053° N, 89.425° W).

The number of roads in the watershed did not change substantially over the

period of the study (see Results section), yet we were interested in whether other

characteristics of the land cover within the watershed had changed over time. To

explore this, we used historical land use/cover data for the watershed that had been

synthesized previously for 1962 and 1995 (Wegener 2001). The original

interpretation of land use/cover was based on ortho-rectified aerial photography. The

United States Department of Agriculture (USDA) captured black-and-white

panchromatic photographs in 1962, and Dane County provided digital orthophotos for

1995. The data for 1962 and 1995 were digitized into a GIS database using ESRI

ArcView (v. 3.1) and on-screen digitizing techniques. Mixed land use/land cover

classes were designated in order to best account for the influence of land-water

interactions.

Page 29: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

18 We also analyzed current land use/cover patterns using data from 2006.

The aerial photographs (1-2m resolution) for 2006 were acquired through the

National Agricultural Imagery Program made available through the USDA Wisconsin

Farm Service Agency. These aerial photographs were digitized in a similar fashion

using ESRI ArcView (v. 9.2) and on-screen digitization techniques. The same land

use/cover designations were used for the 2006 data as were assigned to the 1962

and 1995 photographs.

The land area identified as high-intensity development was assumed to be

completely impervious, while that identified as low-intensity development was

primarily residential. The impervious surface within residential areas was quantified

based on a focused analysis of the proportion of impervious surface (road, driveway,

or rooftop) within land categorized as "low-intensity development". Specifically, ten

sub-set areas each measuring 250 x 250 m were randomly selected within the area

identified as low-intensity development. The exact percent of impervious area (road,

driveway, or rooftop) of each sub-set area was quantified. The mean percent

impervious surface within these ten sub-set areas was then multiplied by the total

surface area classified as low-intensity development to arrive at an estimate of the

total impervious surface within low-intensity development. The high-intensity

development and proportion of the low-intensity development were summed to arrive

at an estimate of the total proportion of the watershed that could be classified as

impervious.

Page 30: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

19 The lake is currently monitored by the Madison Public Health Department

(MPHD) and the University of Wisconsin North Temperate Lakes Long-term

Ecological Research Program (NTL-LTER) on a monthly and quarterly basis,

respectively. MPHD records began in 1962, though consistent monthly monitoring did

not begin until 1973, and data were not collected between 1974 and 1984. NTL-LTER

monitoring of the lake began in 1995 and continues today

(http://lter.limnology.wisc.edu). Therefore, our analysis relies on monthly estimates of

chloride concentrations in 1973 and 1974, 1984 to 1993, and 1995 to 2006. Each

monitoring program collects one or two grab samples from the surface water of the

lake, and chloride samples are analyzed on an ion chromatograph using standard

laboratory methods (USEPA standard method 300.0). During periods when both

MPHD and NTL-LTER records were available for the same month, these estimates

were combined into a single estimate of mean monthly chloride concentration for the

lake. Field replicate samples collected from each monitoring effort indicate that the

natural and analytical variation was low (coefficient of variation < 0.07 in all cases

and < 0.03 in most).

Total road salt applied during each winter has been recorded by the city of

Madison since 1959, and road salt applied specifically to the Lake Wingra watershed

was monitored from 1969 - 1973. While the majority of the watershed is within the

limits of the City of Madison, there are four other municipalities/government agencies

responsible for road management in the Wingra watershed. The City of Madison

maintains 19.3 lane miles (1 lane mile = 1.61 km x number of lanes) (Table 1). The

Page 31: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

20 city limits salting efforts to intersections, hills, bus routes and routes to schools and

hospitals. The Wisconsin Department of Transportation maintains salting on the

Beltline Highway (US Hwy 12/18, 14 and 151, 23.4 lane miles), Dane County

maintains Fish Hatchery Road (6.2 lane miles), the Town of Madison and the Town of

Fitchburg maintain a total of 4.9 lane miles within the watershed (Table 1). Because

road salt applied to the Lake Wingra watershed was not monitored over the entire

period of record, we used three approaches to approximate the application rates in

the watershed based on the observed application rates for the entire city. These

three estimates were then used as candidate predictors in the model fitting procedure

to determine which approximation best explained the observed patterns in the lake.

The first approach (Approach 1) was to use the ratio of the number of miles of

roads in the Lake Wingra watershed compared to road miles for the whole city

(ROADW/ROADM) (Table 2). The ratio was then multiplied by the total road salt

applied in the city (SALTM) to estimate the total salt load for the Wingra watershed for

each winter. The second approach (Approach 2) was to use the number of lane miles

in the Wingra watershed (LaneMilesw) multiplied by the road salt application rate

(AppRateA) and the total number of applications for the winter (TimesAppliedM, as

recorded by the City of Madison) (Table 2). A mean application rate of 68 kg lane-

mile"1 (150 pounds lane-mile"1) was used based on a common set of road salt

application guidelines used across all municipalities and government agencies. In all

cases, we assumed that the number of road miles or lane miles in the watershed has

not changed significantly since the 1970s. This assumption is supported by public

Page 32: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

21 records of roads on the "salt route" during 1972 and 1973 compared to roads

included in the salt routes today (City of Madison Street Division 1973). The final

approach (Approach 3) was to use a regression relationship based on the period

during which there were salt application records for both the City of Madison (SALTM)

and the Lake Wingra watershed (SALTw). The observed relationship between total

road salt applied in Madison and that applied in the Wingra watershed was used to

predict the Wingra-specific application rate in years when Wingra-specific data were

not available. Only five years of data were available for the comparison of Madison to

the Lake Wingra watershed (1969 - 1973), so results of this approximation were

interpreted with care. These three approximations all resulted in an estimate of the

total amount of road salt applied to the watershed during the winter.

These annual estimates were then apportioned into monthly estimates of road

salt application based on three approaches: 1) equal application during winter

months (November - April), 2) application rate proportional to mean observed

monthly application from 1969 - 1973, 3) application rate proportional to the

observed cumulative precipitation during each month. Thus, in the end we had nine

different estimates (3 monthly apportioning methods x 3 estimates of Wingra-specific

loading) of the amount of road salt applied to the Lake Wingra watershed for each

month. Each monthly estimate of road salt application was converted to an estimate

of chloride applied based on the relative molecular weights of sodium and chloride,

which are the primary components of rock salt applied to roadways (0.607 of the total

road salt applied is chloride).

Page 33: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

22 Along with data on chloride concentrations in the lake and road salt applied

in the watershed, we compiled data on meteorological and hydrological variables that

may help to explain chloride dynamics in Lake Wingra. We assessed monthly mean

air temperature, cumulative precipitation between sampling events, estimated

monthly discharge from the lake, estimated CI outflow (based on mean observed CI

concentrations multiplied by estimated hydrologic discharge), days between sampling

events (these were limited to between 21 and 35 days in order to approximate a

monthly time-step for analysis) and the nine different estimates for chloride

application in the watershed. These variables were all used as candidate predictors

in a chloride model at a monthly time step. In addition to data collected/aggregated to

a monthly time step, we collected annual measures of the winter severity index as

calculated and recorded by the Wisconsin Department of Transportation. This index

is based on the incidence of snow events, freezing rain events, total snow amount,

and total storm duration during the winter of interest (Adams, 2009).

All model assessment of monthly data was carried out using the R statistical

and modeling package (http://www.r-project.org/). We first evaluated models to

predict chloride concentrations one time step (month) ahead. These models

correspond to process-error fits as discussed by Hilborn and Mangel (1997). Initial

assessment of candidate predictors was completed using the "regsubsets" function

within the leaps package and initial predictors were selected on the basis of Mallow's

Cp. The top three models from each size class (size class = model set with the same

number of parameters) were then compared using Akaike Information Criterion (AIC).

Page 34: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

The best fitting model was then used to produce a predicted time series of monthly

chloride concentrations based on model parameters and each month's chloride load.

These predicted values were used in model analysis, but were not used in the

simulation exercise.

The monthly data were also fitted to an observation-error method by least

squares (Hilborn and Mangel 1997). The best-fit observation-error model and

parameter values were used in simulation exercises to produce a deterministic

trajectory of future monthly chloride concentrations based on starting conditions (from

2006) and model parameters. We used the observation-error model in the simulation

exercise because we had an estimate of the initial conditions (CI mass in the lake in

2006) and this approach allows prediction of the next time step based on estimated

values rather than observed values (as in the process-error model).

Simulations explored different strategies for road salt management in the Lake

Wingra watershed. We investigated different loading rates ranging from 0 - 60,000

kg month"1 over the next forty years. The initial conditions for the simulations equaled

the current observed mass in the lake (227,910 kg). In each simulation, loading rate

was stochastic with a constant mean and monthly deviations drawn from a normal

distribution with standard deviation equal to the loading rate during the most recent

10 years of observation. Time series plots showed that simulated data became

stationary (constant mean and standard deviation) after approximately 7 years.

Therefore, we allowed each simulation to run for 10 years before recording data in

order to ensure that the model had reached stationarity. We then recorded the

Page 35: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

minimum, maximum and mean chloride concentration in the lake over the next 40

years of simulation.

In a second round of simulations, we examined the response rate of the lake

to changes in road salt application practices. In these simulations, we ran the model

for 25 years at a given load, then decreased the load to half the original load, and ran

the model at the reduced load for an additional 25 years. The onset of lake response

was defined as the point at which the chloride concentration decreased below the

minimum value observed over the 10-year period prior to manipulation. Similarly the

termination of lake response was defined as the point at which the chloride

concentration fell below the maximum CI concentration observed within the final 10

years of the simulation. The response rate was the number of months between the

onset and termination of lake response.

Results

General trends

Since 1962, the amount of road salt applied each winter season has

increased, with the exception of a short period during the early 1970s when the

Madison city council first adopted management policies focused on reducing total

road salt application in the watershed (Figure 1a). These efforts at reduction were

effective until the late 1970s. Concurrently, chloride (CI) concentrations in the lake

have risen over the period of record (Figure 1b). The mean annual CI concentrations

Page 36: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

tended to be more variable than the observed road salt application rate (coefficient

of variation = 0.18 and 0.04, respectively).

Estimates of road salt application rates differed among approaches (Figure 2).

Difference between Approach 1 and 2, though statistically significant (Student's

paired t-test, p < 0.001 in all cases), were not grossly different. However, Approach 3

differed greatly from the other approaches. In all cases, the estimates were not

significantly different based on different methods for allocating total load over the

winter months. All nine estimates of road salt application were used as predictors in

the model, and Approach 1.1 (in which road salt application was proportional to the

number of roads in the Wingra watershed compared to all of Madison, and

application was distributed equally among months) was identified as the best

predictor of patterns observed in the lake.

Despite changes in land use and land cover in the watershed over the past

forty years and accompanied increase in the percent of impervious surface (Figure

3), there has not been a substantial change in the number of road miles maintained

within the Lake Wingra watershed since the 1970's. In 1974 there were

approximately 18.1 lane miles maintained by the city (salt route documented by City

of Madison Street Division 1973), while today there are approximately 19.3 (City of

Madison Engineering Department 2009). Therefore, changes in road salt application

cannot be explained by changes in the total number of roads maintained in the

watershed. Instead, road salt application rates are significantly related to the winter

severity index for Dane County as calculated by the Wisconsin Department of

Page 37: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Transportation (Figure 4) (p = 0.003, R = 0.22). There remains a large amount of

unexplained variation around this relationship. Because the impervious surface

estimates and winter severity index values were only available at decadal or annual

scales, these variables were not included as potential predictors in the monthly

modeling exercise.

Model results

The chloride mass in the lake was best explained by a simple model that

included terms for chloride loading to the watershed and proportional loss from the

lake.

AY = L,_,-bY,_l

Where: AY = change in chloride mass in the lake LM = chloride load during the previous time period YM = chloride mass in lake during previous time period b = chloride loss coefficient

The load term was based on Approach 1.1 for estimating road salt application, in

which the road salt applied in the Lake Wingra watershed was equal to the road salt

applied by the City of Madison multiplied by a correction factor proportional to the

number of roads in Wingra compared to all of Madison. This model was able to

predict the chloride mass in the lake for each month based on observations of

loading and chloride mass from the previous month (Figure 5a and Figure 6a). There

is evidence, however, that this model tends to over predict low values and under

Page 38: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

predict high chloride concentrations. The model was used to predict the entire time

series based on the initial mass in the lake and observed monthly loading (Figure 5c).

Based on this simulation approach, the model was able to represent seasonal and

long-term trends in the data (Figure 5b and Figure 6b).

The above model was then used to examine the effect of different road salt

application practices on the expected equilibrium concentration of CI in Lake Wingra.

The results of the 40-year simulations are summarized as the minimum, mean and

maximum chloride concentration observed in the lake during the 40-year period of

constant loading. The mean CI concentration in the lake increased linearly with an

increase in the mean loading rate used in the simulation (Figure 7). The variation

around CI estimates also increased slightly with an increase in loading rate, as shown

by the increase in the range of simulated values (represented by the error bars in

Figure 7). The mean loading rate (27,442 kg month"1) observed over the last ten

years is represented by a bold line in Figure 7. If this rate of loading continues into

the future, we can expect to witness CI concentrations in the lake over a range of 80

- 150 mg L"1. Observed CI concentrations over the last ten years have ranged from

71.6 - 105.7 mg L"1 (represented by the bold line in Figure 7). Thus, if loading

patterns remain the same, we can expect to see a further increase in CI

concentrations in the lake.

Alternatively, managers may aim to reduce road salt application and thus CI

loading to the lake. A reduction to a constant load equal to 50% of the original loading

rate would result in an eventual 50% decrease in CI concentration in the lake relative

Page 39: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

to the starting concentration (Figure 8). The response rate of the lake varied from

45 to 54 months. Variation in response rate was due to the stochastic nature of the

simulations, and did not differ systematically depending on the initial loading rate or

concentration in the lake. CI is not expected to be retained by the lake, but instead is

removed in proportion to the flushing rate of the lake, which has been estimated to be

~ 0.75 yr"1. The relatively fast flushing rate and the short response time of the model

suggest that Lake Wingra could recover within 5 years from excessive loads, within

the range of chloride loads observed to date.

Discussion

Chloride dynamics in Lake Wingra are primarily driven by the application of

road salt in the watershed. The results of our empirical analysis combined with

modeling efforts suggest that Lake Wingra is sensitive to changes in road salt

application and that the lake responds to changes in road salt management within a

time frame on the order of 45 - 54 months (~ 4 - 5 years). However, loading of

chloride associated with road salt application has begun to affect the shallow and

deep groundwater aquifers in the Madison area, and these subterranean resources

will be much slower to respond to mitigation efforts (Kelly 2008). As groundwater

transport of chloride to the lake increases, the lake may become less responsive to

reductions in road salt application rate to the watershed. Current and short-term

projected mean concentrations for the lake are not above the limit of concern for

health of humans and aquatic organisms (discussed below), yet periodic seasonal

Page 40: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

29 peaks in chloride concentrations do exceed these limits and should be a focus of

further monitoring efforts. A similar model could be applied to other watersheds in

order to explore the repercussions of various road salt management strategies.

Challenges in estimating road salt application

Other studies have shown that the degree of chloride increase in a lake can be

related to the amount of impervious surface in the watershed (Novotny et al. 2008).

Similarly, in the Wingra watershed we see that impervious surface has increased

throughout the period of record. However, in the Wingra watershed, the number of

road-miles has not changed appreciably since the late 1970's. Thus, in Lake Wingra,

we were able to explore the effects of road salt management independent of changes

in road density. It seems that changes in road salt application can be primarily

attributed to changes in management practices that are tied to winter weather

conditions, rather than changes in the length of roads serviced (Figure 4). It remains

possible that an increase in impervious surface may promote more direct delivery of

chloride from the Lake Wingra watershed into the lake. We were not able to directly

examine this idea given the relatively coarse temporal resolution of the data on

impervious surface.

Our model selection procedure consistently identified "Approach 1.1" as the

best metric to represent road salt application; the best long-term metric was based on

a proportion of the number of road miles in the Lake Wingra watershed compared to

the City of Madison with application attributed equally among months of the winter.

Page 41: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Yet, there may certainly be variation among municipal and county managers in

application practices, and temporal variation in how much road salt is applied during

each month of the winter. If application practices are notably different between the

City of Madison, Dane County, the Town of Madison and Fitchburg, then the monthly

estimates of road application may be more variable than those presented here.

Application rates may not differ significantly among management agencies because

all road management agencies rely on the same set of salt application guidelines, but

different management agencies may salt with greater frequency. For example, Dane

County maintains road salt application on a 6-lane highway that serves as the

junction of United States Highways 12, 18, 14 and 151. Given the large volumes of

traffic that use this roadway, we may expect greater application frequency on this

segment of road. It was difficult to assess differences in application rates among

municipalities due to lack of long-term records or event-specific accounts.

Further, road salt estimates used in this study may underestimate the true

application rate within the watershed because we were unable to estimate the

application rate on commercial or residential pavement. Other studies in urban area

have suggested that application to parking lots, sidewalks and driveways may

account for approximately 10 -15% of the total salt application (Howard and Haynes

1993). Overall, our estimates of the magnitude of road salt application in the

watershed should be interpreted as conservative and future modeling efforts would

be improved by a more detailed accounting of when, where and how much road salt

is applied within the watershed.

Page 42: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

31 Imprecise estimates of road salt application may partially explain the

instances of over or under-estimation of chloride concentration that are apparent in

Figure 5b. While the model generally follows the dominant trend of the long-term

data, the model does not do as well in predicting extreme chloride concentrations. It

is possible that during these time periods other factors, like precipitation and

discharge, may be important, though these were not identified as significant

predictors in the model selection procedure. The tendency of the model to under and

over-predict extreme values should be considered in interpreting the results of the

simulations because the range of simulated values may be conservative compared to

the range that can truly be expected.

Accounting for a changing environment

A simple model adequately represents the past and current CI dynamics in

Lake Wingra. However, projections of future CI concentrations are based on the

assumption that the nature of the system will remain the same at higher loading rates

that have not yet been observed in the system. An alternate possibility is that

continued high loading rates could fundamentally change how the system functions,

thus necessitating use of a different model. This will occur if groundwater, which may

constitute up to 35% of the water budget for the lake, becomes further elevated in

chloride.

Historical hydrologic budgets for Lake Wingra indicated roughly equal

contribution from precipitation (31%), surface runoff (34%) and groundwater (35%)

Page 43: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

(Novitzki and Holmstrom 1979). Recent simulated estimates of the hydrologic

contribution of groundwater to the flow of Lake Wingra, however, suggest that

groundwater input to the lake has declined by up to 64% from a historic rate of 3.3 cfs

to a currnet rate of 1.2 cfs (Lathrop et al. 2005) due to municipal groundwater use

and drawdown. Historically, groundwater, surface water and precipitation delivered

freshwater to the system, with the exception of chloride loading via surface runoff

during winter and spring months. However, recent monitoring of the shallow

groundwater aquifer connected to Lake Wingra shows that chloride concentrations

are variable and periodically very high (Figure 9a).

These monitoring data have been collected in conjunction with a groundwater

recharge project within the Lake Wingra watershed. Monitoring wells were installed at

the recharge site, up gradient from the recharge site (to serve as a background

reading), and at two locations down gradient from the recharge site. It is estimated,

based on the groundwater flow paths, that groundwater entering the system at the

recharge site will reach the lake within 20 - 40 years. Increases in chloride

concentrations in the shallow groundwater aquifer could translate into higher

concentrations of chloride entering the lake through groundwater year-round. Under

these conditions, Lake Wingra would take longer to recover following a decrease in

road salt application. Also, the concentrations of chloride in the shallow aquifer would

take much longer to decrease due to the typically low recharge rate and slow flow

rate of groundwater resources (Howard and Haynes 1993, Arnold et al. 2000,

Lindstrom 2005, McGinley 2008).

Page 44: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

33 Lake Wingra is closely connected to the shallow unconfined groundwater

aquifer underlying the watershed (Oakes et al. 1975, Lathrop et al. 2005). Meanwhile,

Madison residents depend on a deep groundwater aquifer (the confined Mount Simon

aquifer), which is used as a municipal drinking water supply. The shallow and deep

groundwater aquifers had historically been characterized as distinct, yet recent

evidence suggests that long-term pumping from the deep aquifer has resulted in

accelerated downward flow of recharge water from the overlying shallow groundwater

aquifer (Borchardt et al. 2007). Increased interaction between shallow and deep

groundwater aquifers raises the possibility that chloride contamination of shallow

groundwater can be conferred to deep groundwater used for drinking water. There is

already evidence of elevated chloride concentrations in monitored drinking water

wells in the Madison area (Figure 9b). The trend of increasing chloride concentrations

in shallow and deep groundwater aquifers, along with the increase in hydrologic

interaction between the two, indicate the need for caution in road salt application in

the region.

Chloride limits for aquatic organisms

Mean model projections are within an acceptable range for health of humans

and aquatic organisms, as current drinking water limits are set at 250 mg L"1 for

human consumption (USEPA 1988). Yet, seasonal peaks in chloride concentrations

in the bottom waters of the lake have been observed to be as high as 420 mg L"1

(Figure 10). A seasonal pattern in chloride concentrations is typical in north-

Page 45: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

34 temperate urban systems because road salt may be flushed off of the landscape in

a pulse and accumulate in the bottom waters of the lake until mixing occurs (Novotny

et al. 2008). During this period, chloride concentrations may exceed the limits for

chronic and acute toxicity for aquatic organisms.

Road salt has been shown to have negative effects on aquatic organisms,

most notably on benthic organisms and larval stages of amphibians (Mayer et al.

2007, Snodgrass et al. 2007, Grapentine et al. 2008). While direct toxicity testing

suggests chronic toxicity limits ranging from 735 - 4681 mg L"1 for specific organisms

in the laboratory, logistic modeling of chronic toxicity indicates that approximately 5%

of aquatic species may be adversely affected by chronic chloride concentrations as

low as 213 mg L"1 (Nagpal et al. 2003). Current USEPA standards take seasonal

variability into account and suggest that aquatic organisms should not be exposed to

a 4-day average concentration of greater than 230 mg L"1 more than once every three

years, and a 1 -hour average chloride concentration exposure should not exceed 860

mg L"1 more than once every three years (USEPA 1988). Based on our calculations,

chronic chloride concentrations in Lake Wingra could reach these harmful levels

(exceeding 213 mg L"1) if loading increases to a mean loading rate of 44,800 kg per

month for a period of 4 years. During the winter of 2007-2008, a winter with high

snowfall in the Madison area, this loading rate was exceeded when a mean of 49,700

kg of chloride was applied each month to the Lake Wingra watershed.

Conclusions

Page 46: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

35 Synoptic studies reveal that road salt application in the United States has

increased steadily since application began in the 1950's (Jackson and Jobbagy 2005,

Kelly 2008). Here we demonstrate a simulation model that explores the response of

an urban lake to changes in road salt management practices. Under current

conditions, Lake Wingra is likely to respond quickly to mitigation efforts (within ~5

years). However, continued loading warrants monitoring of shallow and deep

groundwater resources, due to increasing chloride concentrations, long residence

times, and increasing interaction between shallow and deep aquifers. Similar patterns

in increasing chloride concentrations in surface and groundwaters have been

observed in both urban and pristine areas across northern climates, and similar

modeling approaches may be applied to other systems.

While current mean concentrations of chloride in surface and groundwater in

the Lake Wingra watershed are below the limits of concern for health of human and

aquatic organisms, concentrations periodically exceed these limits. The magnitude

and durations of peaks in chloride concentrations in the lake may not be effectively

documented by current monitoring efforts (based on single grab samples from the

surface of the lake). Response of native biota to periodically high chloride

concentrations is still not clear, and this topic warrants further research.

The approach we present could be a useful tool for managers and decision

makers who are interested in the potential response of surface waters to continued or

increased loading of chloride associated with road salt application. Uncertainty,

especially in the rate of increase and the lag time for recovery of chloride

Page 47: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

concentrations in groundwater, highlight the need for more detailed monitoring of

road salt application and the pathways of interaction between surface and

groundwater resources.

Page 48: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

References 37

Adams, Michael. 2009. Winter severity index. Report to the Maintenance Executive Committee of the Wisconsin Department of Transportation.

Arnold, J. G., R. S. Muttiah, R. Srinivasan, and P. M. Allen. 2000. Regional estimation of base flow and groundwater recharge in the Upper Mississippi river basin. J Hydrol 227:21-40.

Borchardt, M. A., K. R. Bradbury, M. B. Gotkowitz, J. A. Cherry, and B. L. Parker. 2007. Human enteric viruses in groundwater from a confined bedrock aquifer. Environmental Science & Technology 41: 6606-6612.

City of Madison Street Division. 1973. Street Department Deicing Report Lake Wingra Water Shed. Prepared for Madison Rivers and Lakes Commission.

Grapentine, L., Q. Rochfort, and J. Marsalek. 2008. Assessing urban stormwater toxicity: methodology evolution from point observations to longitudinal profiling. Water SciTechnol 57:1375-1381.

Hilborn, R., and M. Mangel. 1997. The ecological detective : confronting models with data. Princeton University Press.

Howard, K. W. F., and J. Haynes. 1993. Urban Geology .3. Groundwater Contamination Due to Road Deicing Chemicals - Salt Balance Implications. Geoscience Canada 20:1-8.

Howard, K. W. F., and H. Maier. 2007. Road de-icing salts as a potential constraint on urban growth in the Greater Toronto Area, Canada. Journal of Contaminant Hydrology 91:146-170.

Interlandi, S. J., and C. S. Crockett. 2003. Recent water quality trends in the Schuylkill River, Pennsylvania, USA: a preliminary assessment of the relative influences of climate, river discharge and suburban development. Water Research 37: 1737-1748.

Jackson, R. B., and E. G. Jobbagy. 2005. From icy roads to salty streams. P Natl Acad Sci USA 102: 14487-14488.

Kaushal, S. S. and others 2005. Increased salinization of fresh water in the northeastern United States. P Natl Acad Sci USA 102:13517-13520.

Page 49: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Kelly, W. R. 2008. Long-term trends in chloride concentrations in shallow aquifers near Chicago. Ground Water 46: 772-781.

Koryak, M., L. J. Stafford, R. J. Reilly, and P. M. Magnuson. 2001. Highway deicing salt runoff events and major ion concentrations along a small urban stream. Journal of Freshwater Ecology 16:125-134.

Lathrop, R. C, K. Bradbury, B. Halverson, K. Potter, and D. Taylor. 2005. Responses to urbanization: Groundwater, stream flow, and lake level responses to urbanization in the Yahara Lakes Basin. LakeLine p. 39-46.

Lindstrom, R. 2006. A system for modelling groundwater contamination in water supply areas: chloride contamination from road de-icing as an example. Nordic Hydrology 37: 41-51.

Mayer, T., Q. Rochfort, U. Borgmann, and W. Snodgrass. 2008. Geochemistry and toxicity of sediment porewater in a salt-impacted urban stormwater detention pond. Environ Pollut 156: 143-151.

McGinley, P. M. 2008. Modeling the influence of land use on groundwater chloride loading to lakes. Lake and Reservoir Management 24:112-121.

Nagpal, N. K., D. A. Levy and D. D. MacDonald. 2003. Ambient water quality guidelines for chloride. The Ministry of Water, Land and Air Protection, British Columbia.

Novitzki, R. P. and B. K. Holmstrom. 1979. Monthly and annual water budgets of Lake Wingra, Madison, Wisconsin, 1972-1977. Water Resources Investigations Report 79-100. Madison, Wl: U.S. Geological Survey, Water Resources Division.

Novotny, E. V., D. Murphy, and H. G. Stefan. 2008. Increase of urban lake salinity by road deicing salt. Sci Total Environ 406:131-144.

Oakes, E. L., G. E. Hendrickson, and E. E. Zuehls. 1975. Hydrology of the Lake Wingra Basin, Dane County, Wisconsin. Water-Resources investigations 17-75. United States Geological Survey.

Schindler, D. W. 2000. Aquatic problems caused by human activities in Banff National Park, Alberta, Canada. Ambio 29: 401-407.

Siver, P. A., A. M. Lott, E. Cash, J. Moss, and L. J. Marsicano. 1999. Century changes in Connecticut, USA, lakes as inferred from siliceous algal remains and their relationships to land-use change. Limnology and Oceanography 44:1928-1935.

Page 50: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

39

Smith, D. W., R. M. Facey, V. Novotny, and D. A. Kuemmel. 1998. Management of winter diffuse pollution from urban areas: Effect of drainage and deicing operations, p. 243-257. In D. E. Newcomb [ed.], Ninth International Conference on Cold Regions Engineering. American Society of Civil Engineers.

Snodgrass, J. W., R. E. Casey, D. Joseph, and J. A. Simon. 2008. Microcosm investigations of stormwater pond sediment toxicity to embryonic and larval amphibians: Variation in sensitivity among species. Environ Pollut 154: 291-297.

Thunqvist, E. L. 2003. Increased chloride concentration in a lake due to deicing salt application. Water Sci Technol 48: 51-59.

United States Environmental Protection Agency (USEPA). 1988. Ambient water quality criteria for chloride. Office of Water, Washington, DC.

Wegener, M. W. 2001. Long-term land use.cover change patterns in the Yahara Lakes Region and their impacts on runoff volume to Lake Mendota. M. S. thesis. Univ. of Wisconsin, Madison.

Page 51: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

40 Table 1. Summary of the number of lane miles maintained by different

municipalities and agencies within the Lake Wingra watershed in 2007.

Municipality/Agency

City of Madison

Wisconsin Department of

Transportation

Dane County

Town of Madison, Town of

Fitchburg

Portions Salted

Intersections; hills; bus,

school and hospital routes

Beltline Highway (US Hwy

12/18, 14, 151)

Fish Hatchery Road

Select streets within the

watershed

Total Lane Miles Managed

19.3

23.4

6.2

4.9

Page 52: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

41

Table 2. Approaches used to estimate the total winter-time load of road salt to the

Lake Wingra watershed. ROADS = the number of road miles (range represents the

range of road miles in Madison between 1973 and 2006), SALT = the total mass of

road salt applied during the winter, LaneMiles = number of lane miles (where a 4-lane

road = 4 x length of road), TimesApplied = Number of times during the winter road

salt was applied (data only available at city level), AppRate = Application rate for road

salt (common guidelines are used across all agencies). Definition of subscripts: W =

APPROACH

ROADSw 1- ROADSM

M

2. LaneMi les w * T imesApp l ied M * AppRateA

3. S A L T w = b 0 + b 1 * S A L T M

CONDITIONS

ROADSw = 19.3 mi ROADSM (range = 512 - 758 mi) SALTM (range =1379-16277 MT)

LaneMilesw = 57 mi TimeAppliedM (range = 16 - 36) AppRateA = 68 kg mi"1

Observed for 1969 - 1973 only SALTw (range = 238 - 459 MT) SALTM (range = 2895 - 6071 MT) |3o, |3i = coefficients used to predict SALTw in other years

Lake Wingra watershed, M = City of Madison, A = All agencies.

Page 53: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

„#—•*,

' l -

nJ 0 >,

1 -^ » • •*-*

sal

T3 cc o cr

o o "fr

o o CO

200

o o

o H

(a)

• -•

• • • • •

o o

D)

F ° •S- °° c o 03 i_

"c CD O c o o O

o CO

o

o CM

1960 1970 1980 1990 2000

(b)

. /I 1' •

t'i w 2010

1 1

i >

• • •

r i i i i i i

1960 1970 1980 1990 2000 2010

Figure 1. (a) Total road salt applied in the Lake Wingra watershed during each winter, (b) Mean annual chloride concentration observed in Lake Wingra based on monthly sample collection following 1973. Prior to 1973, estimates are based on 1 - 6 observations of chloride concentrations in the lake within the year. Error bars represent the standard deviation around the annual mean (only shown for years with more than 2 samples).

Page 54: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

43

1^^ 1

x: c o E

XL

• o .92 "5. Q. 05

T\ w

o o o o CM

O o o

150

000

o o

o o o o m

o H

1 2 3

Approach 1

1 2 3

Approach 2

1 2 3

Approach 3

Figure 2. Estimates of road salt application rates based on three approaches for estimating the total load for the winter (Approach 1, Approach 2, Approach 3) and on three methods for allocating road salt application among months of the winter (methods are designated as 1, 2, 3). The approaches are fully described in Table 1. Solid lines represent the median of the data, notches represent a 95% confidence interval, the boxes represent the inter-quartile range, the whiskers and dots represent the full range in the data, with dots signifying potential outliers. Each Approach is significantly different from the other (Student's t-test, p-values <0.001 in all cases), but within each Approach there was no significant difference among methods for allocating among months.

Page 55: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

1962

•'J1|"~V

J:/

u

40

35

30

25

« 20 o 3

0)

a. E

15

10

5

^ 3

' 2 8

A ^ s

X_J

Agriculture

Forest

High Intensity Dev.

Low Intensity Dev.

L—) Cemetery

1995 D

Wetland

• Golf Course

Open Water

J • Grassland

2006

-,-J

1962 1995 2006

Figure 3. Changes in land cover in the Lake Wingra watershed over 44 years. Though the number of road miles in the watershed has not changed significantly in the period between 1995 and 2006, there has been an increase in impervious surface in the watershed.

Page 56: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

10 20 30 40 50 60

Winter severity index

Figure 4. Significant positive relationship between the winter severity index and the mean winter-time chloride load in the watershed (p = 0.003, R2 = 0.22). The line represents the model fit to the data. Winter severity index values were available from 1993-2008.

Page 57: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

46

O)

.* <D JC TO

_C

(0 to ro E o

o

000

o 8 O »o CM

o o 8

- Observed o Predict one step

°5-

° o 0 c

o o

* rP

-V Jb -

r

" o o

• ^ - " c l •» 8° 0 - o "b

- o o o

>̂ .

> ° - _ °

IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIMIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIMIIIIIIIIIIIIIIIIIIIIIIMIIMIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIMIMIIIIIIllllllllllllllllllllllllllllllll 1973-03-27 1984-10-01 1986-03-03 1987-10-01 1989-04-03 1990-11-07 1992-08-01 1997-05-01 2001-06-26 2003-05-29 2006-03-27

<0

- Observed o Predict whole series

o

"*"

-

• " O

- - - "A »°° -fln oas^ o » - „ o - ° ^ *T* o " -* O w o w n «• O O " D O — °

° ° o ° J> ^ - - o o ° o ° ° o

_ - - - ° o°

"rf*

-

o

™ o

m j . OO

"*. - o o

° *•" ° oo

oo o« ° » o . J 3 . ° . o - J > 0 o # o ° .

„o J- . - o~ o ° - * ° w

- o — _ • o . ^ • -o » ° •

" o —

" iiiiiiiiiiiiMiiiiiiiiiiiiiiiiiiiiiiiiMiiiiiiiiiiiMiiiiiiiiiiiMiiiiiiiiiniiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiniiiiiiiHiiiiiiiiiiiiiiiniiiiiiiiiiiiiiiiiiiiiiiiiiiiiiii

1973-03-27 1984-10-01 1986-03*3 1987-10-01 1989-04-03 1990-11-07 1992-08-01 1997-05-01 2001-06-26 2003-05-29 2006-03-27

Q . Q . CO

O

I I I I I I M I I I I I I I I I I I I I I I I I I

1973-03-27 1984-10-01 iii iMiiii i i i iHiiiniii i i i i i i i i i i i i i i i i i i i i i i i i i inniiii i i i i i i i i i i i i i i i i i i i i iniii i i i iHiiiHiii i i i i i i iMiiNiiii iniii i ininiii i i i i i i i i i i iniii i— 1986-03-03 1987-10-01 1989-04-03 1990-11-07 1992-08-01 1997-05-01 2001-06-26 2003-OS-29 2006-03-27

Year-Month-Day

Figure 5. Chloride mass in the lake that was observed and predicted by the best fitting model. Predictions were made for each month based on the previous observed chloride mass (process-error model) (a), and predictions were simulated based on the initial conditions and model parameters (observation-error model) (b). The amount of CI applied in each month served as input to the model (c). Note the x-axis is not continuous.

Page 58: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Pre

dict

ed [C

O in

lake

(m

gL

1)

Pre

dict

ed [C

I] in

lake

(m

g L 1

)

0 20

40

60

80

10

0 12

0 14

0 0

20

40

60

80

100

120

140

o

-

o

o

CJ o

o o o

K5 o

O

I I

I J\ \ \

\

\ •

"

I I

\ f>

;fe

^jmk'

«>

^Sg

jl^

^ 'fw

*£*"

?tl»

TB

djfc

*

WJ

^ l>\

--.

... c%

-i>

*"

\ r?

\

. \ \

\ • •

*^

o

o o

o o

o

O

O

Page 59: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Figure 6. Observed and predicted values for the model. The line represents a 1:1 relationship. Panel (a) presents the fit one-step ahead (process-error model). Panel (b) presents fit to whole time series (observation-error model).

E, CD O)

c

CD

CO

c g * • » CO

c 0 o c o o o

o CM

O o CM

O in

o o

o

o H

i i r 0 10000 20000 30000 40000 50000 60000

CI Load (kg month"1)

Figure 7. Mean chloride concentration expected in Lake Wingra following forty years of loading at different loading rates. The error bars represent the minimum and maximum simulated concentration based on stochastic simulations of a constant load. The bold line represents the range in CI concentrations observed over the past ten years at a mean winter loading rate of 27442 kg month"1.

Page 60: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

II

*"'i / ',

\/\f x/\j v V . vV\

V S A A A A A A / V ^

• ; •• I I

%(*.

A /

— Csirrent load x 2 - - - Current load x 1.5 + Current load

Current load x 0.5

v v V v v v _

I ':

0 50 100 150 200 250

Time step of simulation (month)

Figure 8. Effects of a 50% reduction in chloride loading rate. The baseline loading rate was set to four different levels: 2 times the current load, 1.5 times the current load, the current load, and 0.5 times the current load. The simulated reduction in load was implemented at time step 110.

Page 61: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

50

10/2/04 4/20/05 11/6/05 5/25/06 12/11/06 6/29/07 1/15/08 8/2/08 2/18/09

Date

100

90

80

70

60

50

40 -j

30

20

10

(b) • UW6

DUW14

OUW17

XUW23

ft x

• D

X

1 O * x

o o X X* o

Xx

QDDHt«» • o • «**

• •

1970 1975 1980 1985 1990 1995 2000 2005 2010

Year

Figure 9. Chloride concentrations in wells that tap the shallow (a) groundwater aquifer and deep (b) groundwater aquifer (the Mount Simon aquifer). The horizontal line in panel a represents the EPA limit for chronic exposure for aquatic organisms. The deep groundwater aquifer is used as a drinking water source by the City of Madison. Note the difference in scales on the axes.

Page 62: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

0.00

0.50

1.00

.§. 150 1 sz \ % 2-00 ] Q !

2.50 j

3.00 -I

I 3.50 +-

51

100 200 300 400

Chloride concentration (mg L-1) 500

slh Figure 10. Chloride profile in Lake Wingra on March 18 , 2008. Depth represents the depth from the surface of the lake.

Page 63: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

52

CHAPTER 2

Phosphorus sources and demand during summer in a eutrophic lake

by

Amy M. Kamarainen1,3, Rachel M. Penczykowski1,2, Matthew C. Van de Bogert1, Paul

C. Hanson1 and Stephen R. Carpenter1

1 Center for Limnology, University of Wisconsin, Madison, Wisconsin 53706, USA

2 Current address: School of Biology, Georgia Institute of Technology, Atlanta,

Georgia 30332, USA

3 Corresponding author: [email protected]

Keywords: ecosystem metabolism; entrainment; respiration; production; C:P ratio;

phytoplankton bloom.

Status: Kamarainen, A. M., R. M. Penczykowski, M. C. Van de Bogert, P. C. Hanson,

and S. R. Carpenter. 2009. Phosphorus sources and demand in a eutrophic lake.

Aquatic Sciences 71: 214-227.

Page 64: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

53 Abstract

In pelagic systems, phytoplankton biomass may remain abundant or near

equilibrium while concentrations of the limiting nutrient are below detection. In

eutrophic lakes, it has been thought that episodic algal blooms are due to mixing

events that break down this equilibrium by adding nutrients to the mixed layer.

Alternatively, rapid rates of biotic recycling among primary producers and

heterotrophic consumers could maintain high phytoplankton biomass, yet the

recycling process has been difficult to observe in situ. Here we use free-water oxygen

measurements and an associated metabolic model to infer rates of phosphorus (P)

uptake and biotic mineralization in the epilimnion of a eutrophic lake. The rates of

uptake and mineralization were compared to "external" sources of P such as loading

and entrainment. Also, model results were assessed using sensitivity analysis. We

found that the majority of phytoplankton P demand during the period of low P

availability could be accounted for by biotic mineralization, but that it was important to

consider the effects of entrainment in order to account fully for P uptake. These

general results were relatively insensitive to model parameterization, though the

relative C:P ratio of material taken up versus mineralized was an important

consideration. This study integrates modeling and measurement tools that monitor

ecosystem processes at finer temporal resolution than has previously been possible,

complementing other studies that use experimental incubation and elemental tracers.

Extension of this approach could enhance models that aim to integrate biological and

Page 65: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

54 physical processes in assessment of water quality and prediction of phytoplankton

biomass.

Page 66: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Introduction

Phytoplankton in lake ecosystems may remain abundant under apparent

nutrient limitation despite loss by sinking, grazing, natural mortality and bacterial- and

viral-mediated mortality (Juday et al., 1927; De Pinto et al., 1986; Caraco et al.,

1992). This paradox is particularly puzzling in eutrophic lakes where strong summer

stratification and nutrient limitation would seem to favor phytoplankton loss over

production and growth, yet phytoplankton blooms may occur during the period when

phosphorus concentrations are below detection (De Pinto et al., 1986). A number of

mechanisms may explain the provisioning of phosphorus (P) to nutrient-limited

primary producers including biotic recycling, loading from external sources and

entrainment of P-rich metalimnetic waters.

Biotic recycling of organic P compounds by heterotrophic zooplankton and

microbes is highly variable, but field and laboratory research suggests that much of

the P needed for phytoplankton production may be supplied by heterotrophic

mineralization (Barlow and Bishop, 1965; Cole et al., 1988; Sterner, 1989; Poister et

al., 1994) and that mineralization may be the dominant supply of nutrients during

bloom formation (De Pinto et al., 1986). Excretion of inorganic P by zooplankton

alone may account for 4-239% of the P demand of primary producers, and this

percentage varies among months (Gulati et al., 1995; Vanni, 2002). However, other

analyses focused on rates of primary production that are corrected for loss of

phytoplankton argue that biotic recycling is not sufficient to explain observed rates of

production during the summer (Caraco et al., 1992). Thus, multiple lines of evidence

Page 67: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

suggest a variable and important role of heterotrophic consumers in P recycling,

yet this process has been difficult to observe and quantify in lake ecosystems

(Lehman, 1980).

External loading and entrainment are dominant sources of nutrients to lakes at

an annual time scale, yet their importance during the period of summer stratification

and low nutrient availability is not clear. External loading is often thought to be

unimportant during summer months due to lower seasonal flows into lakes. Similarly,

entrainment, or transport of P from deep water, has been noted to be low due to

strong stratification (Lean et al., 1987). Recent work, however, suggests that the flux

of nutrients across the thermocline may be an important driver of epilimnetic

metabolism (Maclntyre et al., 2002; Staehr and Sand-Jensen, 2007). Another study

shows that external P inputs fuel new production (Caraco et al., 1992). Notably,

external loading and entrainment may be most important following storms (Maclntyre

et al., 2006), but in these cases the event is difficult to monitor and the fate of nutrient

influx can be difficult to trace. As such, the relationship between episodic storms,

movement of P-rich waters, and blooms of phytoplankton is noisy and difficult to

interpret (Soranno et al., 1997; Maclntyre et al., 2002).

The pathways of P supply are transient and occur at temporal and spatial

scales that have been challenging to observe. This challenge has been met using

light/dark bottle incubations with 14C as a proxy for metabolic P demand, 32P

incubations as a measure of P uptake rates, and experimental isolation of

heterotrophic organisms to estimate rates of mineralization (Ryther, 1956; Hargrave

Page 68: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

and Geen, 1968; Lean et al., 1983; Lean et al., 1987; Sterner, 1986; Vanni, 2002).

These approaches involve extended incubation times during which it is difficult to

maintain representative environmental conditions. Additionally, there is unresolved

debate about the relationship between rates of 14C fixation and gross versus net

primary production (Peterson, 1980; Carignan et al., 2000). The challenges and

limitations associated with these techniques are widely recognized (Bender et al.,

1987; Smith and Prairie, 2004; Staehr and Sand-Jensen, 2007).

Here we explore a complementary free-water approach to estimate P uptake

and mineralization. We derived estimates of in situ P uptake and heterotrophic

mineralization using high-resolution oxygen data and an associated metabolic model.

We then compared metabolically-inferred P uptake and mineralization rates to

sources of P such as external loading and entrainment. Empirical estimates of

uptake, mineralization, loading and entrainment were used to calibrate a model of P

demand and supply and we used this model to explore the theoretical bounds of

biotic and abiotic P transformation within the epilimnion of a eutrophic lake. Our goal

was to examine the dynamic behavior of P demand and supply processes to

elucidate the relative importance of biotic recycling and abiotic sources of P like

external loading and entrainment, particularly during the period of low P availability.

Materials and Procedures

This study has four main components: empirical measurements of oxygen,

carbon and P; integration of these empirical estimates into a whole-ecosystem

Page 69: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

metabolism model to estimate P uptake and mineralization rates; comparison of

metabolically-derived uptake rates to potential sources of P such as in situ

heterotrophic mineralization, external P loading and entrainment of P from the

metalimnion of the lake; and sensitivity analysis of model results to changes in our

assumptions. Estimated uptake rates served as a proxy for P demand, and

throughout the text we use "uptake" and "demand" interchangeably.

All data were collected during the summer of 2007 on Lake Mendota, a

eutrophic lake in Madison, Wisconsin (43°06' N, 89°25' W, 39.1 km2 surface area,

12.3 m mean depth). Lake Mendota stratifies during the summer and soluble reactive

P (SRP) typically falls below the limits of analytical detection in the epilimnion (North

Temperate Lakes Long Term Ecological Research program

(http://lter.limnology.wisc.edu)). Because SRP is often used as a proxy for the

concentration of bioavailable P in the system, we targeted our assessment on this

component of the P budget. We focused sampling efforts on the epilimnion during the

period of strong stratification and low concentrations of SRP (28 June 2007 until 8

September 2007) in order to explore how phytoplankton P demand during this period

may be met.

Field and Laboratory Methods

During the summer of 2007 an automated buoy equipped with temperature

and oxygen probes was deployed at a central location in Lake Mendota. A TempLine

(Apprise Technologies, Inc.) thermistor array recorded temperature every minute at

Page 70: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

0.5 - 1.0 m intervals from 0 - 20 m depth. Dissolved oxygen concentrations were

monitored using a D-Opto (Zebra-Tech LTD, New Zealand) optical dissolved oxygen

probe that was suspended from the buoy at 2 m. The dissolved oxygen probe was

compared weekly to manual measurements and corrected, assuming linear drift. The

high-resolution temperature data were averaged over 24 hour periods and used to

identify the maximum depth of the mixed layer for each day. Diel mixed layer depth

was around 8 m for the season, though on a number of days microstratification was

apparent within the first 2 or 3 meters of the surface. In order to avoid incorrectly

identifying shifts in depth of the mixed layer due to microstratification, we defined the

depth of the mixed layer as the final depth (starting from lake surface) at which the

temperature was within 1.5 °C of the mean temperature of the first 5 m.

Water samples were collected at least weekly in acid-washed polyethylene

bottles at 1 - 4 m intervals throughout the water column using a peristaltic pump for

determination of total P (TP) and soluble reactive P (SRP) concentrations. In the

field, SRP samples were filtered through a 0.45-u,m polycarbonate etched filter using

in-line filtration, and stored on ice. TP samples were preserved using Optima HCI,

while SRP samples were refrigerated and analyzed within 24 h. SRP samples were

analyzed colorimetrically using the ascorbic acid method (American Public Health

Association (APHA), 1995). TP samples were analyzed using a Technicon Auto

Analyzer following persulfate digestion (APHA, 1995). All statistical analyses and

model iterations were run using the mean P concentration of samples from the

epilimnion.

Page 71: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

60 Particulate carbon and P samples were collected from triplicate integrated

epilimnetic samples. These samples were collected at least weekly from mid-June to

mid-October at depths determined by concurrent temperature profiles. All samples

were filtered onto glass fiber filters and stored for carbon or P analysis. Because

there is no evidence that the C:P ratio of phytoplankton should differ markedly from

that of the whole water seston (Healey and Hendzel, 1980; Sterner and Elser, 2002),

particularly during the summer when phytoplankton dominate seston composition

(Hecky et al., 1993; Elser et al., 1995), the C:P ratios for uptake and mineralization

are derived from the C:P ratio of the entire seston. Carbon values used in the C:P

ratio were based on determination of ash-free dry mass (AFDM). Three replicates of

each sample were filtered onto pre-weighed glass fiber filters (Proweigh GF/F, 47

mm, 1.5 urn pore size). The filters with sample residue were placed in a drying oven

for at least 48 h, transferred to a desiccator for at least 48 h, and weighed. The filters

and samples were then combusted at 550 ° C for 4 h, returned to the drying oven and

desiccator, and weighed again. Combusted mass was subtracted from dry mass to

determine AFDM. Seston carbon concentration was inferred as 48% of the AFDM

(Round, 1965; Fogg, 1975). Whole seston particulate P concentrations used as the

denominator in the C:P ratio were determined using the method described by

Lampman et al. (2001). Integrated epilimnetic samples were filtered onto pre-

combusted glass fiber filters and frozen until analysis. Samples, standards, and

blanks were placed in 25 ml acid-washed serum vials and digested by adding 20.0 ml

of 1% low N potassium persulfate (Fisher Scientific ID #P282-500). The vials were

Page 72: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

61 sealed with aluminum rings crimped over butyl rubber septa and autoclaved for 2.5

h at 120 ° C. The concentration of P in the liquid portion of each sample was

analyzed colorimetrically using the ascorbic acid method and all samples were

corrected for P content of the filter and of the digestion reagent (APHA, 1995).

Water samples for the determination of chlorophyll a concentration were

collected twice per week at 2 m using a peristaltic pump or Van Dorn sampler. Water

was collected in dark 3.5 L bottles and stored on ice in the field. Within three hours

of collection, samples were mixed well, filtered under low light conditions onto glass

fiber filters (Whatman GF/F, 47 mm, 0.7 u.m pore size), and frozen until analysis.

Chlorophyll a samples were extracted with methanol and analyzed fluorometrically on

a Turner TD-700 fluorometer. The final concentrations were corrected for

phaeopigments (Holm-Hansen and Riemann, 1978; Arar et al., 1997).

Metabolic Inference of Phosphorus Uptake and Mineralization

Estimates of gross primary production (GPP) were derived as follows from

free-water oxygen measurements collected during buoy deployment. For each 1-

minute measurement interval, t, we calculated net ecosystem production (NEPt) from

the measured change in oxygen concentration, A0 2 (mmol 0 2 m"3 min"1), and

atmospheric exchange, Dt (mmol 0 2 m"3 min"1). To express metabolism in areal units

(mmol m"2 min"1), we multiplied by the depth of the mixed layer, z (m).

(Eq. 1) NEP, = GPP,-Rt = (A0 2 -D t ) *z

Page 73: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Atmospheric exchange can be positive or negative. We use positive values to

indicate addition of O2 to the lake and negative values for removal. Atmospheric

exchange was estimated as

k([02]SAT,-[02]t) (Eq. 2) D, = '

[O2] SAT, t (mmol m"3) is the aqueous concentration of oxygen if the lake were in

equilibrium with the atmosphere and was calculated from water temperature using

the empirical equation of Weiss (1970). [02]t (mmol m"3) is the measured

concentration of dissolved oxygen in measurement interval t. We calculated the gas

piston velocity, k (m min"1), using estimates of k6oo (m min"1) as a function of wind

speed (Cole and Caraco, 1998; Cole et al., 2000) and the water temperature

dependent Schmidt number, Sc, (Wanninkhof, 1992).

(Eq.3) t -my" 5

We measured wind speed every minute at 2 m above the water surface and

converted these measurements to values at 10 m height assuming a neutrally stable

boundary layer and the empirical equation of Smith (1985) to calculate values of k6oo

Page 74: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

(m min ). In order to express atmospheric exchange in volumetric units, we

divided by the depth of the mixed layer, z (m).

We used the values of NEP for each measurement interval (NEPt) to obtain

daily estimates of NEP, GPP and ecosystem respiration (R) (Eq. 1). During darkness,

the NEPt values are attributable solely to respiration. Therefore, we summed the

nighttime NEPt values and divided by the period of darkness to get the rate of

ecosystem R at night. We followed the convention of assuming daytime R is equal to

nighttime R and averaged the R rates obtained for the night preceding and night after

each daylight period to estimate R for each day. Daily values of GPP were calculated

by summing the interval measurement of NEPt for the daylight hours and adding

daytime R. We aggregated daily estimates of GPP and R to a weekly scale following

the convention established by Cole et al. (2000) and justified by Staehr and Sand-

Jensen (2007). Also, the weekly scale was most fitting for inference because P

samples were collected at a weekly time step.

We used estimates of GPP and ecosystem respiration (Rtot) to estimate P

demand and mineralization (assuming photosynthetic and respiratory quotients of

1.0). The assimilation of P into algal cells should occur at a rate proportional to the

net primary production (NPP) observed in the system. NPP is equivalent to GPP

(mmol O2 m"2 d"1) corrected for the amount of oxygen used in autotrophic respiration

(Ra) (mmol 02 m"2 d"1). Thus, to determine the rate of NPP and convert the oxygen-

based estimate to carbon we used the following equation:

Page 75: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

64

(Eq. 4) NPP = 0.375* (GPP-R a )

Where:

NPP = net primary production (mmol C m"2 d"1)

Ra = autotrophic respiration (mmol O2 m"2 d"1)

0.375 = mass ratio of C to O2

Autotrophic respiration is difficult to determine, but has been quantified in a number of

studies with values generally ranging from 35 - 60% of total community respiration

(del Giorgio and Peters, 1993; Duarte and Cebrian, 1996, Dodds and Cole, 2007). As

a nominal value we assumed that Ra would be equal to 50% of total community

respiration.

In order to infer the P uptake rate (Puptake), we assumed that over the short

period of the study the in situ C:P ratio of phytoplankton serves as an adequate

estimate of the C:P ratio at which inorganic materials are assimilated during

photosynthesis. Thus, from estimates of NPP (mmol C m"2 d"1) and seston molar C:P,

we calculated rates of P uptake based on the following:

(Eq.5) Puptake=NPP*(l/(C:P))

Mineralization of P (Pmin), a measure of biotic recycling, was assumed to occur

under conditions of heterotrophic metabolic equilibrium (i.e. growth ~ loss). Bacterial

Page 76: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

biomass is relatively constant during this period of the summer in Lake Mendota

(epilimnetic bacterial biomass ranged from 130.3 - 177.8 mg C m"2 in 1979 and 84.7

- 167.0 mg C m"2 in 1980; Pedros-Alio and Brock, 1982) and it is likely that overall

heterotrophic growth rates are approximately balanced by loss. Under a scenario of

zero net heterotrophic growth, we infer that the C:P of mineralization is equivalent to

the C:P ratio of the substrate. Therefore, the C:P ratio of mineralization was assumed

to be equal to the C:P ratio of whole epilimnetic seston. This value was multiplied by

the proportion of total respiration due to heterotrophs (Rh) (mmol C m"2 d"1) according

to the following equation:

(Eq.6) Pmin=Rh*(l/(C:P))

Where:

Rh = Rtot - Ra

Estimating non-metabolic sources of phosphorus

In addition to quantifying the metabolic processing of P, we also assessed

alternate SRP sources to the lake during the period when SRP was below detection.

Estimates of external load were based on an approach used previously (Lathrop et

al., 1998; Carpenter and Lathrop, 2008). There are four streams and two storm water

inlets that enter Lake Mendota. Two streams (Pheasant Branch and Yahara River)

and one storm water inflow (Spring Harbor) are continuously monitored for hydrologic

and chemical inputs into the lake by the United States Geological Survey (USGS).

Page 77: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

66 Loading was determined for these three inlets and these data were used to infer

loading from other inlets based on previous estimates of the relative load entering the

lake from each source (Lathrop et al., 1998). Baseline P loads were measured as

input of total P, but data were also available to determine the approximate proportion

of P entering the lake in the soluble form. For most inflows (Six Mile Creek, Pheasant

Branch, Spring Creek, Spring Harbor and Willow Creek), the proportion of P entering

as SRP was 24-52% of TP and did not depend on discharge. However, for the largest

inflow, the Yahara River, the proportion entering the lake as SRP differed

systematically based on flow rates (SRP = 0.045 (± 0.024) x TP when discharge <

5.7 m3 s"1; SRP = 0.56 (± 0.028) x TP when discharge > 5.7 m3 s"1) due to a wide

river estuary immediately upstream of the lake. Therefore, when discharge > 5.7 m3

s"1, the "high flow" proportion (0.523 x TP) of SRP was attributed to the input, while

during "low flow" SRP input was equal to 0.186 x TP, based on flow-weighed mean

SRP:TP ratio of all inflows. This rate was used to convert estimates of TP loading

from all inflows to areal estimates of SRP loading. Similarly, we were interested in

accounting for any significant loss of SRP to downstream systems through the outlet

of the lake. The hydrologic outflow was multiplied by the mean SRP concentration of

the epilimnion during the periods when SRP was detectable, however the total mass

of SRP lost through this pathway was negligible (mean outflow < 0.0001 mg SRP m"2

d"1).

Entrainment of SRP into the epilimnion from lower strata of the lake was also

considered. Entrainment events were identified using the high-resolution temperature

Page 78: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

profiles, and entrainment was defined as an increase in the maximum depth of the

mixed layer by 0.5 m or more. When an additional mass of water was incorporated

into the epilimnion we quantified the associated SRP flux based on the volume of

water incorporated and the mean SRP concentration of that water mass. This flux

includes SRP derived from hypolimnetic and benthic remineralization.

We also considered the input of P from rain water using total weekly

precipitation, as measured at a weather station associated with the Dane County

Truax airport, multiplied by a TP concentration of 0.032 mg P L"1, an average used in

previous budget calculations for the lake (Lathrop, 1979; Soranno et al., 1997). The

mean estimate we derived (0.19 mg P m"2 d"1 ± 0.26) is likely an overestimate of the

input of SRP because SRP makes up only a portion of total P input through

precipitation.

Release of SRP from epilimnetic sediments in contact with the mixed layer

was also considered as a source. The mean weekly depth of the mixed layer was

used to calculate the area of the sediment surface that was in contact with the

epilimnion, and this value was multiplied by an average SRP release rate for Lake

Mendota of 2.4 mg P m"2 d"1 (Stauffer, 1987; Soranno et al., 1997). These estimates

amounted to a mean release rate of 0.72 mg P m2 d"1 (± 0.04) for the sediments in

contact with the epilimnion, as the sediment area in contact with the mixed layer is

approximately 30% of total lake sediments. Average inputs from sediment release

and from precipitation were consistently low, hence we focused on external loading

and entrainment, which can be episodic and important during summer months.

Page 79: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

68

Sensitivity Analysis

In order to assess all variables on a common temporal scale, the mean weekly

values of each input variable (GPP, Rtot, C:P, external load, entrainment) were cast

into the model to compare the relative magnitude of SRP uptake and mineralization,

and assess their magnitude in comparison to other sources of P (Fig. 1a). Rather

than rely solely on mean values to summarize the trends in P demand and supply,

however, we explored the feasible range of values that served as input to the model

through sensitivity analysis. The full suite of mean weekly values observed for each

variable was incorporated into the analysis through a bootstrapping procedure (Fig.

1 b). The model was run for 2000 iterations by drawing values randomly with

replacement from the observed dataset. Because GPP and Rtot were positively

correlated (Pearson's correlation analysis, r = 0.83, p = 0.002), these two variables

were drawn concurrently from the dataset, meanwhile C:P values were not correlated

with GPP and Rtot, therefore the C:P ratio for uptake (C:Pup) and mineralization

(C:Pmin) were drawn independently and randomly from the pool of observed weekly

mean seston C:P ratios.

While we lack evidence that the C:P ratio of natural phytoplankton populations

differs significantly from that of overall seston, one might think of scenarios of high

detrital content or low phytoplankton biomass in which such a difference could occur.

Also, while difference in the C:P ratio among components of the seston may be

imperceptible based on current seston separation techniques, it is certainly possible

Page 80: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

69 that different trophic groups (bacteria versus zooplankton) are differentially using

seston components in metabolic processes. Thus, there is likely to be an imbalance

between the C:P ratio taken up by phytoplankton and the C:P ratio of substrate that is

metabolized and subsequently mineralized during respiration. We explored the

sensitivity of the model results to the influence of an imbalance between the C:P ratio

of uptake and mineralization. Such exploration was based on the observed range in

weekly C:P ratios (180.1 - 483.8), which was comparable to sestonic C:P ratios that

had been previously documented for temperate lakes (range = 122-441) (Hecky et

al., 1993; Dobberfuhl and Elser, 2000)

Sensitivity analyses were used to explore the full range of values that may

feasibly be observed for the photosynthetic quotient (PQ), respiratory quotient (RQ)

and Ra. PQ and RQ vary depending on the biochemical composition of the molecules

produced or broken down during the metabolic process. Many researchers assume a

baseline value of 1.0 for both PQ and RQ in aquatic ecosystems (del Giorgio and

Peters, 1993; Hanson et al. 2003), and empirical estimates support this assumption

(Bender et al., 1987), yet reported values may range from 0.8 - 1.2 (del Giorgio and

Peters, 1993). We explored uncertainty in the quotients by running the model using a

range of PQ and RQ values from 0.8 - 1.2.

Another variable with potentially high uncertainty was the percent of total

community respiration (Rtot) that could be attributed to autotrophs (Ra) and

heterotrophs (Rh). Total respiration tends to increase with lake trophy, and respiration

by autotrophs becomes an increasing part of total respiration as lakes become more

Page 81: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

eutrophic (del Giorgio and Peters, 1993; Biddanda et al., 2001; Roberts and

Howarth, 2006). A literature review by del Giorgio and Peters (1993) suggests that

autotrophic respiration may contribute 35% in oligotrophic systems and over 60% in

eutrophic systems. As a conservative estimate of the range expected in natural lakes

we used Ra values that ranged from 30-70% of total community respiration in the

sensitivity analysis.

Results

P concentrations followed a seasonal trend typical for Lake Mendota (Stauffer,

1987; North Temperate Lakes Long Term Ecological Research program

(http://lter.limnology.wisc.edu)). Mean epilimnetic SRP concentrations declined

through the spring and reached the limit of analytical detection by week 27 (1 July

2007 - 7 July 2007) (Fig. 2a). During spring and early summer (week 16-25, 15

April 2007 - 23 June 2007), SRP and TP concentrations were correlated (r = 0.97, p

« 0.001). In contrast, TP was closely related to the particulate P concentration after

week 25 (24 June 2007 - 8 September 2007) (r = 0.75, p = 0.008), while SRP fell

below detection. Following week 25 (24 June 2007), the majority of P mass in the

epilimnion was bound within the particulate pool (TP = 0.032 ± 0.002 mg P L"\

Particulate P = 0.025 ± 0.002 mg P L"1). At the same time, the SRP concentration in

the bottom waters of the lake increased (Fig. 2b). Our analysis was focused on the

period of the summer between weeks 26 and 34 (24 June 2007 - 25 August 2007)

when the lake was strongly stratified and SRP concentrations were below the limit of

Page 82: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

71 detection (<0.003 mg L"1). Despite a decline in biologically available P,

phytoplankton biomass remained relatively high. Chlorophyll a concentration reached

a peak in week 29 (15 July 2007 - 21 July 2007) (52.9 ±1.9 ug L"1), and thereafter

centered around a mean of 16.8 ± 5.8 ug L'1 for the remainder of the summer (Fig.

2c). The molar C:P ratio of seston in the epilimnion was generally high early in the

summer (275.1 - 370.9) (20 June 2008 - 4 August 2008) and was lower during the

later part of the summer (108.7 - 221.2) (5 August 2008 - 6 October 2008) (Fig. 2d).

Gross primary production (GPP) ranged from 0.86 - 2.30, with an average of 1.52 g

C m"2 d"1. Total community respiration (Rtot) ranged from 0.32 - 1.56, with an average

of 1.08g Cm"2d"1.

TN:TP ratios in the epilimnion ranging from 25.2 to 84.7 (on molar basis)

suggested that phosphorus was limiting. Similarly, DIN:SRP ratios in the metalimnion

(8 - 10 m) during the period of interest ranged from 15.4 - 30.7, suggesting that

water entering the epilimnion through entrainment would be relatively rich in

bioavailable N (based on Redfield's N:P ratio of 16:1). Thus, P was considered the

limiting nutrient during our study.

During the period that P was below the detection limit (24 June 2007 - 25

August 2007), weekly rates of SRP uptake were relatively constant (Fig. 3). While

uptake rates were relatively low in weeks 28 - 30, rates were not significantly

different among weeks 28 - 33. In most weeks, uptake exceeded the amount of P

supplied by any single source, with the exception of weeks 28 and 33 (8 July 2007 -

14 July 2007 and 12 August 2007 - 18 August 2007). Rates of P supply via

Page 83: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

72 mineralization were more variable than rates of uptake during the period of

interest. Mineralization ranged from 23 - 109% of uptake, thus meeting an average of

57% of documented phytoplankton P demand. Mineralization represented a roughly

consistent source of SRP to phytoplankton throughout the summer (mean = 4.5 ±

0.94 mg SRP m"2 d"1), while entrainment (mean = 2.59 ±1.41 mg SRP m"2 d"1)

occurred during irregular pulses. During such events, the SRP flux via entrainment

was comparable to observed rates of mineralization. P loading from external sources

was consistently low during this mid-summer period and did not satisfy a significant

portion of phytoplankton P demand (0.28 ± 0.03 mg SRP m"2 d"1, 3.5% of demand).

These trends taken together showed that on average the mean daily uptake rate

(7.83 ± 0.66 mg SRP m"2 d"1) was met by the sum of documented sources (7.38 ±

2.38 mg SRP m"2 d"1) (t-test, p = 0.39); note that these two groups are independent

over this period so a t-test was appropriate (Fig. 4). However, no single source of P

was sufficient to explain the variation in P uptake. The mean rate of mineralization

was not significantly different than that of entrainment (t-test, p = 0.41).

Sensitivity Analysis

Model iterations using combinations of low, nominal and high values for PQ,

RQ and Ra demonstrated that the rate of uptake and the sum of all sources of P were

not significantly different (by t-test) in any of the model scenarios (Fig. 5). The large

error bars around the sum of sources in figure 5, representing the standard deviation

around the mean, can be partly attributed to variation in estimates of entrainment and

Page 84: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

73 external loading. External loading and entrainment estimates were independent of

the metabolic parameters of the model, thus the effect of changes in model

parameters may be best explored by comparing the differences between P uptake

and mineralization.

In comparing P uptake to mineralization alone, high PQ/low RQ scenarios

showed that uptake was significantly greater than mineralization in all cases (Fig. 5c).

Given nominal PQ/RQ values, we found the relationship between uptake and

mineralization was mediated by the value of Ra. When autotrophic respiration (Ra)

was low, uptake was more likely to be balanced by mineralization (Fig. 5b). When PQ

was low (0.8) and RQ (1.0) was high, however, model results showed that

mineralization could feasibly meet the demands of phytoplankton P uptake under any

Ra scenario. Model results were most clearly affected by changes in the relative

values of PQ and RQ, while model results were less sensitive to changes in the

proportion of total community respiration attributed to autotrophs (Ra).

Given the difficulty in measuring C:P ratios of different sestonic components,

we explored the sensitivity of model output to the C:P ratio used in calculating uptake

and mineralization rates. We found the relative C:P ratio (C:Pup/C:Pmin) was a critical

factor in determining the magnitude of the difference between uptake and

mineralization rates inferred from the model (Fig. 6). Results of all model iterations

(2000) are presented in figure 6 as the difference between uptake and mineralization.

For this sensitivity analysis, GPP, Rtot, external loading and entrainment were

integrated into the model as a dependent dataset, while the C:P ratios for uptake and

Page 85: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

mineralization were randomly and independently chosen from the entire pool of

observed C:P ratios of seston. Thus, the results presented in figure 6 demonstrate

the effect of our assumption that the C:P ratios of uptake and mineralization are

equal. As long as C:Pup was smaller than C:Pmin (C:PUp/C:Pmin < 1.0) the rate of P

uptake exceeded that of P supplied through mineralization. We found that if C:Pmin

and C:Pup are significantly different in nature (which is possible, but hard to measure),

then these differences could affect our estimates of the relative balance between

i uptake a n d Imin-

We were primarily interested in how P demand was met during periods of low

P availability, yet we also had data to investigate patterns in P uptake and supply

during the later period of the summer when P was available. An expanded time

series showed the influence of a large storm with significant precipitation (and runoff,

not shown) that occurred between week 33 and week 34 (Fig. 7). This storm event

resulted in input of SRP from both external loading and entrainment (Fig. 8). Likely in

response, mean uptake rates increased during week 35 (Fig. 8). During weeks 37

and 38, progressive deepening of the mixed layer at the end of the summer season

resulted in significant entrainment and subsequent increases in the observed uptake

and mineralization rates during weeks 38, 39 and 40.

When these results were averaged over the extended 14-week period,

mineralization and uptake became statistically indistinguishable (t-test, p = 0.29) (Fig.

9). This was largely due to higher variation in uptake and mineralization represented

in the larger data set (CV increased from 0.46 to 0.64 and from 0.55 to 0.91 for

Page 86: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

uptake and mineralization, respectively). The sum of sources, though not

statistically different from uptake, indicated that P was supplied in excess of what was

needed by primary producers. The availability of P at the end of this period was also

apparent in the epilimnetic P trends presented in figure 2a.

Discussion

We explored the mechanics and magnitude of P demand and supply using a

combination of empirical data and modeling. Results indicated that the amount of P

used by primary producers during the period of low P availability was predominantly

supplied via biotic mineralization. P demand, however, cannot be met by

mineralization alone and the relative balance among P sources (mineralization,

external loading and entrainment) varied over the summer. Entrainment was also an

important pathway of P transport, and the episodic input of P occurred during a

period of higher than average wind speeds (week 27). While these results were

generally insensitive to model conditions (i.e. independent of changes in Ra, PQ and

RQ), the balance between P uptake and mineralization was mediated by the relative

C:P ratios of uptake and mineralization. Thus, our results, which showed that

patterns of P demand and supply were dictated by the relative C:P ratio of producers

and consumers, support the importance of stoichiometric relationships in foodweb

interactions and ecosystem processes (Elser et al., 1988; Elser et al., 1995;

Dobberfuhl and Elser, 2000).

Page 87: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Our finding that entrainment was necessary to account for observed uptake

suggests that entrainment can be important not only in annual budgets but also as a

contributor to meeting P demand at a weekly scale during the stratified P-deficient

season. Thus, part of the variation observed in patterns of gross primary production

(GPP) and net ecosystem production (NEP) may be attributed to exchange of

material across the thermocline (Staehr and Sand-Jensen, 2007). Also, entrainment

occurred during a period of higher than average wind speeds (5.3 m s"1 during week

27, compared to June-September average of 4.5 m s"1), but this entrainment event

was independent of other storm indicators (precipitation, changes in solar flux). Such

observations may help explain why the relationship between phytoplankton blooms

and storm events are exceedingly noisy (Soranno et al., 1997).

Interestingly, peaks in phytoplankton biomass occurred approximately one

week after significant P fluxes from entrainment and external loading (weeks 29, 35

and 38). The initial chlorophyll a peak (week 29), however, occurred soon after

depleted P conditions were apparent. It is possible that the bloom we witnessed

during week 29 was due to growth of the phytoplankton population that began under

SRP replete conditions, possibly sustained by luxury uptake. Given our observed

average uptake rate of 10 mg m"2 d"1 during week 26 and 27 and a mean

mineralization of approximately 5mg m"2 d"1, we can come to a net P demand of

approximately 5mg m"2 d"1, which translates to a daily uptake rate of 196 kg d"1.

Meanwhile, the observed epilimnetic SRP concentration in week 26 was 0.0046 mg

L"1, equaling an estimated 955 kg of SRP in the epilimnion. Given the conditions in

Page 88: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

week 26, we would expect the SRP available in the epilimnion to have been taken

up within approximately 5 days. This uptake may also have included luxury uptake

under nutrient replete conditions. Given that phytoplankton may sustain three or four

cell-doublings without taking up additional phosphorus (Reynolds 2006), a growth

rate of 0.2 d"1 (or a doubling time of ~3.5 days) may allow phytoplankton to persist for

an additional 14 days on luxury uptake. Based on the above assumptions, we can

conclude that a bloom beginning in week 26 may have grown and persisted over a

maximum of 19 days without a new source of phosphorus.

Later peaks in chlorophyll a concentration (weeks 35 and 38), though,

demonstrate that entrainment and external loading can induce peaks in

phytoplankton biomass that are accompanied by increases in the rates of uptake. It

seems that mineralization generally provides sufficient P to sustain production during

periods of low P availability, but that an external pulse is required to induce higher-

than-average chlorophyll a concentrations. Our results corroborate other work

showing that physical processes, including entrainment, are linked to external drivers

(meteorological variation, evaporative cooling, temperature of inflow) in complex

ways and can play an important role in nutrient cycling and ecosystem processes

(Maclntyre et al., 2002; Fragoso et al., 2008). Our approach based on free-water

metabolism estimates can be used to assess the relative importance of biotic

recycling, entrainment and external loading among systems, and to reduce

uncertainties in the relationship between P supplies and phytoplankton biomass

during summer months.

Page 89: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

We explored uncertainties in the model using alternative sets of model

conditions (i.e. a range in values for Ra, PQ and RQ); in all cases inferred P uptake

was equal to the sum of sources to the epilimnion during the period of low P

availability. Under scenarios of unbalanced photosynthetic and respiratory quotients

(PQ = 0.8, RQ = 1.2, all Ra) or low Ra (PQ = 1.0, RQ = 1.0, Ra = 0.3), mineralization

alone was sufficient to account for total P demand, these model scenarios, however,

are relatively unlikely to occur. Theoretical calculations and empirical observations

suggest that PQ values are likely to be > 1.0 (Ryther, 1956; Williams and Robertson,

1991). The PQ was estimated to be 1.34 for a typical algal cell comprised of 40%

protein, 40% carbohydrate, 15% lipid, and 5% nucleic acid (Williams and Robertson,

1991). Meanwhile, best estimates of RQ, while variable, tend to be < 1.0 (Hutchinson

and Edmondson, 1957; Lampert and Bohrer, 1984). Thus the conditions represented

by the PQ = 0.8 and RQ = 1.2 scenario are unlikely in aquatic ecosystems, while the

conditions of our PQ = 1.2 and RQ = 0.8 scenario are more plausible. The results of

our sensitivity analysis demonstrate that different values of Ra may affect the

magnitude of uptake and mineralization, yet the difference between uptake and

mineralization is relatively insensitive to changes in the proportion of total community

respiration attributed to autotrophs (Ra) versus heterotrophs (Rh). When PQ and RQ

are unbalanced, the pattern of uptake and mineralization will be dictated by the

PQ.RQ relationship rather than the balance between autotrophic and heterotrophic

respiration.

Page 90: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

79 While model output was generally insensitive to changes in PQ, RQ, and

Ra, the relative balance between uptake and mineralization was sensitive to the

stoichiometric relationship between primary producers and organisms contributing to

heterotrophic mineralization. Transfer of carbon and nutrients through foodwebs is

mediated by stoichiometric relationships among foodweb components and trophic

patterns may be structured by the stoichiometry of primary producers at the base of

the foodweb (Sterner and Elser, 2002). There is variability in the C:N:P ratios of

seston in lakes, and many organisms at the base of the foodweb have some plasticity

in maintenance of tissue C:N:P ratios (Hecky et al., 1993; Sterner and Elser, 2002;

Diehl et al., 2005). While we were not able to empirically distinguish among sestonic

foodweb components (bacteria, phytoplankton, micrograzers), we were able to

explore the influence of variation in C:P ratios within the modeling framework.

Our model results help constrain expectations of the patterns of P supply and

demand likely in natural systems. Given the relationship presented in figure 6 and

previously published stoichiometric relationships, we argue that P demand is likely to

outweigh P supplied via mineralization alone. There is no conclusive evidence that

the C:P ratio of seston, the component available as substrate for mineralization,

should differ from the C:P ratio of phytoplankton, the component that represents the

ratio of uptake (Healey and Hendzel, 1980; Sterner and Elser, 2002). Yet, there is

evidence that mineralization of organic matter by zooplankton can exacerbate P

limitation due to excretion at relatively high C:P ratios (Elser et al., 1988; Sterner,

1990). Also, measured C:P ratios of bacteria and zooplankton tend to be lower than

Page 91: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

80 those documented for whole-water seston (Goldman et al., 1987; Fagerbakke et

al., 1996; Dobberfuhl and Elser, 2000; Hochstadter, 2000). As such, retention of P in

bacterial and zooplankton pools likely results in C:P ratios of mineralization that are

greater than the C:P ratios of primary producers (Elser et al., 1995), and thus the

relative ratio of C:Pup:C:Pmin is likely to be < 1.0. Under such conditions, our model

results indicate that natural lake ecosystems are likely to cluster within the upper, left-

hand portion of figure 6, and it may not be possible to explain the mid-summer rate of

primary production by P supplied through mineralization alone.

Bacteria have been highlighted as dominant decomposers and mineralizers of

organic material, yet they may also compete with phytoplankton for uptake of

bioavailable P (Rigler, 1956; Currie and Kalff, 1984; Cotner and Wetzel, 1992). If

bacteria are significant contributors to rates of heterotrophic respiration and these

bacterial populations are net consumers of P, then our model would underestimate

the true ecosystem demand for bioavailable P, and overestimate the rate of

mineralization. While we do not have data to address this possibility directly, previous

estimates of bacterial populations in Lake Mendota during late summer demonstrate

relative consistency in bacterial biomass (Pedros-Alio and Brock, 1982). Given this

consistency, it is likely that growth in the bacterial population is closely mirrored by

loss and that the net population growth and rate of P uptake are close to zero. Also,

Cotner and Wetzel (1992) suggest that P taken up by bacteria does not represent net

uptake because they generally lack the capacity for P storage, thus bacterial uptake

is comparable to the amount lost from internal pools. In light of the small influence of

Page 92: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

81 heterotrophic versus autotrophic respiration on model results, we do not expect

changes in our assumptions about the influence of bacteria on uptake and

mineralization to have large effects on the results of the analysis.

Our approach and findings contribute to understanding of the relationship

between primary production and nutrient cycling in aquatic ecosystems and

complement other studies conducted using experimental incubation and elemental

tracers. Using an approach that relies on free-water estimates, we can explore

ecosystem processes in natural ecosystems, instead of under laboratory, light/dark

bottle, or microcosm conditions. Our work integrates modeling approaches and

monitoring data enhanced by tools for high-frequency data acquisition that can allow

monitoring of processes at finer temporal resolution than has previously been

possible. The assumptions of our approach are different from those inherent in

previous analyses and therefore, this approach could be used to corroborate or

question results from other modeling, experimental and observational approaches.

Also, our work highlights an application for the emerging area of research related to

whole-ecosystem metabolism (Hanson et al., 2003; Staehr and Sand-Jensen, 2007).

As our estimates of whole-ecosystem metabolism improve, we stand to gain a better

understanding of ecosystem processes like nutrient cycling. Further application and

refinement of this approach could also enhance models that assess water quality and

predict phytoplankton biomass through integrated representation of biological and

physical processes (Hamilton and Schladow, 1997; Fragoso et al., 2008).

Page 93: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

References

American Public Health Association, American Water Works Association, and Water Pollution Control Federation, 1995. Standard methods for the examination of water and wastewater, 19 ed. American Public Health Association.

Arar, E. J., G. B. Collins, and United States. Environmental Protection Agency. Office of Water, 1997. Method 445.0 in vitro determination of chlorophyll a and pheophytin a in marine and freshwater algae by fluorescence, Revision 1.2. ed. United States Environmental Protection Agency Office of Research and Development National Exposure Research Laboratory.

Barlow, J. P. and J. W. Bishop, 1965. Phosphate regeneration by zooplankton in Cayuga Lake. Limnology and Oceanography 10:15-24.

Bender, M., K. Grande, K. Johnson, J. Marra, P. J. LeB. Williams, J. Sieburth, M. Pilson, C. Langdon, G. Hitchcock, J. Orchardo, C. Hunt, P. Donaghay and K. Heinemann, 1987. A comparison of 4 methods for determining planktonic community production. Limnology and Oceanography 32:1085-1098.

Biddanda, B., M. Ogdahi, and J. Cotner, 2001. Dominance of bacterial metabolism in oligotrophic relative to eutrophic waters. Limnology and Oceanography 46: 730-739.

Caraco, N. F., J. J. Cole, and G. E. Likens, 1992. New and recycled primary production in an oligotrophic lake - insights for summer phosphorus dynamics. Limnology and Oceanography 37: 590-602.

Carignan, R., D. Planas, and C. Vis, 2000. Planktonic production and respiration in oligotrophic shield lakes. Limnology and Oceanography 45:189-199.

Carpenter, S. R. and R. C. Lathrop, 2008. Probabilistic estimate of a threshold for eutrophication. Ecosystems 11: 601-613.

Cole, J. J., S. Findlay, and M. L. Pace, 1988. Bacterial production in fresh and saltwater ecosystems - a cross-system overview. Marine Ecology Progress Series 43: 1-10.

Cole, J. J. and N. F. Caraco, 1998. Atmospheric exchange of carbon dioxide in a low-wind oligotrophic lake measured by the addition of SF6. Limnology and Oceanography 43: 647-656.

Page 94: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Cole, J. J., M. L. Pace, S. R. Carpenter, and J. F. Kitchell, 2000. Persistence of net heterotrophy in lakes during nutrient addition and food web manipulations. Limnology and Oceanography 45:1718-1730.

Cotner, J. B. and R. G. Wetzel, 1992. Uptake of dissolved inorganic and organic phosphorus-compounds by phytoplankton and bacterioplankton. Limnology and Oceanography 37: 232-243.

Currie, D. J. and J. Kalff, 1984. A comparison of the abilities of fresh-water algae and bacteria to acquire and retain phosphorus. Limnology and Oceanography 29: 298-310.

Del Giorgio, P. A. and R. H. Peters, 1993. Balance between phytoplankton production and plankton respiration in lakes. Canadian Journal of Fisheries and Aquatic Sciences 50: 282-289.

De Pinto, J. V., T. C. Young, J. S. Bonner, and P. W. Rodgers, 1986. Microbial recycling of phytoplankton phosphorus. Canadian Journal of Fisheries and Aquatic Sciences 43: 336-342.

Diehl, S., S. Berger, and R. Wohrl, 2005. Flexible nutrient stoichiometry mediates environmental influences, on phytoplankton and its resources. Ecology 86: 2931-2945.

Dobberfuhl, D. R. and J. J. Elser, 2000. Elemental stoichiometry of lower food web components in arctic and temperate lakes. Journal of Plankton Research 22: 1341-1354.

Dodds, W. K. and J. J. Cole, 2007. Expanding the concept of trophic state in aquatic ecosystems: It's not just the autotrophs. Aquatic Sciences 69: 427-439.

Duarte, C. M. and J. Cebrian, 1996. The fate of marine autotrophic production. Limnology and Oceanography 41:1758-1766.

Elser, J. J., M. M. Elser, N. A. Mackay, and S. R. Carpenter, 1988. Zooplankton-mediated transitions between N-limited and P-limited algal growth. Limnology and Oceanography 33:1-14.

Elser, J. J., T. H. Chrzanowski, R. W. Sterner, J. H. Schampel, and D. K. Foster, 1995. Elemental ratios and the uptake and release of nutrients by phytoplankton and bacteria in 3 lakes of the Canadian Shield. Microbial Ecology 29:145-162.

Page 95: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Fagerbakke, K. M., M. Heldal, and S. Norland, 1996. Content of carbon, nitrogen, oxygen, sulfur and phosphorus in native aquatic and cultured bacteria. Aquatic Microbial Ecology 10: 15-27.

Fogg, G. E„ 1975. Algal cultures and phytoplankton ecology, 2d ed. University of Wisconsin Press.

Fragoso Jr., C. R., D. M. L Motta Marques, W. Collischonn, C. E. M. Tucci, and E. H. Van Nes, 2008. Modelling spatial heterogeneity of phytoplankton in Lake Mangueira, a large shallow subtropical lake in South Brazil. Ecological Modelling (in press)

Goldman, J. C, D. A. Caron, and M. R. Dennett, 1987. Regulation of gross growth efficiency and ammonium regeneration in bacteria by substrate C-N ratio. Limnology and Oceanography 32:1239-1252.

Gulati, R. D., C. P. Martinez, and K. Siewertsen, 1995. Zooplankton as a compound mineralising and synthesizing system: Phosphorus excretion. Hydrobiologia 315: 25-37.

Hamilton, D. P. and S. G. Schladow, 1997. Prediction of water quality in lakes and reservoirs .1 . Model description. Ecological Modelling 96: 91-110.

Hanson, P. C, D. L. Bade, S. R. Carpenter, and T. K. Kratz, 2003. Lake metabolism: Relationships with dissolved organic carbon and phosphorus. Limnology and Oceanography 48:1112-1119.

Hargrave, B. T. and G. H. Geen, 1968. Phosphorus excretion by zooplankton. Limnology and Oceanography 13: 332-342.

Healey, F. P. and L. L Hendzel, 1980. Physiological indicators of nutrient deficiency in lake phytoplankton. Canadian Journal of Fisheries and Aquatic Sciences 37: 442-453.

Hecky, R. E., P. Campbell, and L. L. Hendzel, 1993. The stoichiometry of carbon, nitrogen, and phosphorus in particulate matter of lakes and oceans. Limnology and Oceanography 38: 709-724.

Hochstadter, S., 2000. Seasonal changes of C : P ratios of seston, bacteria, phytoplankton and zooplankton in a deep, mesotrophic lake. Freshwater Biology 45: 372-372.

Page 96: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Holm-Hansen, O. and B. Riemann, 1978. Chlorophyll a determination -Improvements in methodology. Oikos 30: 438-447.

Hutchinson, G. E. and Y. H. Edmondson, 1957. A treatise on limnology. Wiley.

Juday, C, E. A. Birge, G. I. Kemmerer, and R. J. Robinson, 1927. Phosphorus content of lake waters of northeastern Wisconsin. Transactions of the Wisconsin Academy of Sciences 23: 233-248.

Lampert, W. and R. Bohrer, 1984. Effect of food availability on the respiratory quotient of Daphnia magna. Comparative Biochemistry and Physiology a-Physiology 78: 221-223.

Lampman, G. G., N. F. Caraco, and J. J. Cole, 2001. A method for the measurement of particulate C and P on the same filtered sample. Marine Ecology Progress Series 217: 59-65.

Lathrop, R. C, 1979. Appendix H: Lake Management, p. H1 - H73. Dane County Regional Planning Commission.

Lathrop, R. C, S. R. Carpenter, C. A. Stow, P. A. Soranno, and J. C. Panuska, 1998. Phosphorus loading reductions needed to control blue-green algal blooms in Lake Mendota. Canadian Journal of Fisheries and Aquatic Sciences 55:1169-1178.

Lean, D. R. S., A. P. Abbott, M. N. Charlton, and S. S. Rao, 1983. Seasonal phosphate demand for Lake Erie plankton. Journal of Great Lakes Research 9:83-91.

Lean, D. R. S., A. A. Abbott, and F. R. Pick, 1987. Phosphorus deficiency of Lake Ontario plankton. Canadian Journal of Fisheries and Aquatic Sciences 44: 2069-2076.

Lehman, J. T., 1980. Release and cycling of nutrients between planktonic algae and herbivores. Limnology and Oceanography 25: 620-632.

Macintyre, S., J. R. Romero, and G. W. Kling, 2002. Spatial-temporal variability in surface layer deepening and lateral advection in an embayment of Lake Victoria, East Africa. Limnology and Oceanography 47: 656-671.

Macintyre, S., J. O. Sickman, S. A. Goldthwait, and G. W. Kling, 2006. Physical pathways of nutrient supply in a small, ultraoligotrophic arctic lake during summer stratification. Limnology and Oceanography 51:1107-1124.

Page 97: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

86

Pedros-Alio, C. and T. D. Brock, 1982. Assessing biomass and production of bacteria in eutrophic Lake Mendota, Wisconsin. Applied Environmental Microbiology 44:203-218.

Peterson, B. J., 1980. Aquatic primary productivity and the C-14-C02 method - a history of the productivity problem. Annual Review of Ecology and Systematics 11:359-385.

Poister, D., D. E. Armstrong, and J. P. Hurley, 1994. A 6-Yr record of nutrient element sedimentation and recycling in 3 north temperate lakes. Canadian Journal of Fisheries and Aquatic Sciences 51: 2457-2466.

Reynolds, C. S., 2006. Ecology of phytoplankton. Cambridge University Press, Cambridge, UK.

Rigler, F. H., 1956. A tracer study of the phosphorus cycle in lake water. Ecology 37: 550-562.

Roberts, B. J. and R. W. Howarth, 2006. Nutrient and light availability regulate the relative contribution of autotrophs and heterotrophs to respiration in freshwater pelagic ecosystems. Limnology and Oceanography 51: 288-298.

Round, F. E., 1965. The biology of the Algae. Edward Arnold.

Ryther, J. H., 1956. The measurement of primary production. Limnology and Oceanography 1:72-84.

Smith, S. V., 1985. Physical, chemical and biological characteristics of C02 gas flux across the air water interface. Plant, Cell and Environment 8: 387-398.

Smith, E. M. and Y. T. Prairie, 2004. Bacterial metabolism and growth efficiency in lakes: The importance of phosphorus availability. Limnology and Oceanography 49:137-147.

Soranno, P. A., 1997. Factors affecting the timing of surface scums and epilimnetic blooms of blue-green algae in a eutrophic lake. Canadian Journal of Fisheries and Aquatic Sciences 54:1965-1975.

Staehr, P. A. and K. Sand-Jensen, 2007. Temporal dynamics and regulation of lake metabolism. Limnology and Oceanography 52:108-120.

Page 98: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Stauffer, R. E., 1987. Vertical nutrient transport and its effects on epilimnetic phosphorus in 4 calcareous lakes. Hydrobiologia 154: 87-102.

Sterner, R. W., 1986. Herbivores direct and indirect effects on algal populations. Science 231: 605-607.

Sterner, R. W., 1989. Resource competition during seasonal succession toward dominance by cyanobacteria. Ecology 70: 229-245.

Sterner, R. W., 1990. The ratio of nitrogen to phosphorus resupplied by herbivores -Zooplankton and the algal competitive arena. American Naturalist 136: 209-229.

Sterner, R. W. and J. J. Elser, 2002. Ecological stoichiometry : the biology of elements from molecules to the biosphere. Princeton University Press.

Vanni, M. J., 2002. Nutrient cycling by animals in freshwater ecosystems. Annual Review of Ecology and Systematics 33: 341 -370.

Wanninkhof, R., 1992. Relationship between wind-speed and gas-exchange over the ocean. Journal of Geophysical Research-Oceans 97: 7373-7382.

Weiss, R. F., 1970. Solubility of nitrogen, oxygen and argon in water and seawater. Deep-Sea Research 17: 721-735.

Williams, P. J. L. and J. E. Robertson, 1991. Overall planktonic oxygen and carbon-dioxide metabolisms - the problem of reconciling observations and calculations of photosynthetic quotients. Journal of Plankton Research 13: S153-S169.

Page 99: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

88 Acknowledgments

The authors would like to draw particulate attention to the assistance of Mark Kemfert

in the field, to James Thoyre and Mark Lochner for assistance with laboratory

analyses, and to Luke Winslow and Dave Balsiger for assistance with data access.

Discussions with Dick Lathrop, Jonathan Cole and Paul del Giorgio significantly

improved the manuscript. Financial support was provided by the National Science

Foundation graduate student fellowship program (AMK), the National Science

Foundation Research Experience for Undergraduates (RMP). Field work, buoy

instrumentation, and data processing were supported by the North Temperate Lakes

Long Term Ecological Research Program, the Global Lakes Environmental

Observatory Network, and the University of Wisconsin - Madison Department of

Atmospheric and Oceanic Sciences.

Page 100: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

89

(a) BASELINE APPROACH

02, Temperature, Wind speed

' ' Metabolism Model

1'

Estimates of: GPP. Rto,

> '

Estimates of: C:P

i -

Uptake/Mineralization Model

Assumes: PQ=RQ=1.0 Ra=0.5

Weekly estimates of:

Uptake and Mineralization

(b) SENSITIVITY ANALYSIS

Weekly estimates of:

GPP, R(ot, C:P, External loading, Entrainment

BOOTSTRAP (Draw with replacement)

Uptake/Mineralization Model

Scenarios: Ra = 0.3, 0.5, 0.7 PQ,RQ = 0.8,1.0, 1.2

Model estimates of:

Uptake, Mineralization, and Sum of Sources

Figure 1. Flow chart depicting the steps taken in the baseline model run and in the sensitivity analysis.

Page 101: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

0.14

0.12

£, c o

0.10

0.08

c <u o c o o

0.06

0.04

0.02

0.00

(a)

o | Soluble Reactive P • I Particulate P

o o 0

•!

o •

I poooooc^ .OOP

, - p 15 20 2 5 l _ _ . 3 0 _ _ l 35 40

May Jun Jul Aug Sep Oct

45

0.5

0.4

E,

o

8 c o

o 0. tr CO

0.3

0.2

0.1

0.0

IP) I 1

o 12m 16 m

o 0»

• •

o o

15 20 25 I 30__J 35 40 45

May Jun Jul Aug Sep Oct

60 -

50 -

^̂ ZJ 4 0 -

O) 3

"ST 3° " xz Q. 20 -

2 O 1 0 -

0 -

(c)

• •

'

F = I

i i i i i ! i T * l± I

I* I

I I I I

1 1

=-=^1

j I • *

J* 1* * • • • •

1 ' '

600

15 20 2 5 l _ _ 3 0 i 35 40 45

Week of Year May Jun Jul Aug Sep Oct

Week of Year May Jun Jul Aug Sep Oct

Figure 2. Summer trends in epilimnetic phosphorus concentration (a), selected hypolimnetic soluble reactive phosphorus (SRP) concentration (b), phytoplankton biomass (as measured by chlorophyll a concentration) (c) and seston C:P ratio (d). Values are mean estimates for each week, error bars represent one standard deviation. The dotted box delineates the focal period for this study.

Page 102: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

91

i "O

CM I

Q.

E, x D

0=

c CD CD

«fr _

CM J

O J

00 H

co H

-* H

CM

o -•

D Uptake E3 Mineralization M External load • Entrainment

2a_ dm— I M ®-U

4\

2m.

27 28 29 30 31 32 33 Week of Year

JULY AUGUST

Figure 3. Mean weekly estimate of rates of soluble reactive phosphorus (SRP) uptake, mineralization, external loading and entrainment for the period when SRP was below the limit of detection during summer 2007. Error bars represent one standard error around the mean and account for the daily variation in estimates of GPP and Rtot.

Page 103: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

92 o _

oo

I

a. a: CO CD

X 3

5=

C CO CD

<o H

* H

CM - H

o - J

Uptake 7.83

Sum Sources 7.38

Mineralization 4.5

Source

VMM External Load

0.28 Entrainment

2.59

Figure 4. Overall mean rate of phosphorus uptake compared to estimated magnitudes of possible sources of phosphorus. Sum Sources represents the sum of all documented sources of SRP input during the P-deficient stratified period. The components of Sum Sources (mineralization, external load, and entrainment) are shown to the right. Error bars represent one standard error around the mean (n = 7 weeks). Input of SRP from precipitation and epilimnetic sediment release were less than 1 mg SRP m"2 d"1 and are not shown here.

Page 104: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

93

16

14

12

10

8

6

4

2

0

(a)PQ

I 0.8

I RQ1.2

i' • Uptake • Mineralization

II 0.3 0.5 0.7

(b)PQ1.0, RQ1.0 • Uptake • Mineralization

0.3 0.5 0.7

Figure 5. Results of sensitivity analysis based on alternate values (0.8, 1.0, 1.2) for the photosynthetic quotient (PQ) and the respiratory quotient (RQ) as well as alternate values for the percent of total respiration that is autotrophic (Ra = 0.3, 0.5, 0.7). Error bars represent the standard deviation around the mean. Asterisks (*) indicate a significant difference between mean uptake and mineralization (t-test, a = 0.05). The difference between uptake and the sum of sources was not significant in any of the scenarios tested (t-test, a = 0.05).

0.3 0.5 0.7

R ;

Page 105: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

94

i 1 1 1 r 0.5 1.0 1.5 2.0 2.5

C:Pup/C:Pmin

Figure 6. The relationship between the relative difference between uptake and mineralization and the ratio of C:P for uptake relative to mineralization. When the relative C:P ratio is close to one, as was the case in the baseline portion of our results, the difference between uptake and mineralization is slightly positive. When the C:P ratio of mineralization is less than C:P of uptake (C:Pup/C:Pmin > 1 -0), then the difference between uptake and mineralization approaches a mean of zero. However, when C:P of mineralization is greater than uptake (C:Pup/C:Pmin < 1.0), the difference between uptake and mineralization is always positive.

Page 106: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

95

T 1 1 1 1 r 22 24 26 28 30 32 34 36 38 40

i " ' i " i r

22 24 26 28 30 32 34 36 38 40

800

600

O 400 H

T 1 1 1 1 r 1 r

22 24 26 28 30 32 34 36 38 40

Figure 7. Meteorological data for the summer of 2007. All data were monitored at the Atmospheric and Oceanic Sciences Building at University of Wisconsin - Madison, located 0.8 km south of Lake Mendota. Data available at: http://rig.ssec.wisc.edu/.

Week

Page 107: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

CM

Q. a: CO

o>

X

Q.

c CO 0) ? 4

D Puptake E Pmineral @ SRPIoad • SRPentrain

27 28 29 30 31 32 33 34

Week

35 36 37 38 39 40

Figure 8. Mean weekly estimate of rates of soluble reactive phosphorus (SRP) uptake, mineralization, external loading and entrainment for an extended period during summer 2007 (n=14 weeks). Error bars represent one standard error around the mean and account for the daily variation in estimates of GPP and Rtot.

Page 108: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

r* i

•a M 1

E Q. a: m

1 flux

(m

g I

lean

P

o

in

o

W

^ ^

Uptake 13.11

Sum Sources 15.37

Mineralization 9.62

Source

Load 1.97

Entrainment 3.78

Figure 9. Overall mean rate of phosphorus uptake compared to estimated magnitudes of possible sources of phosphorus. Sum Sources represents the sum of all documented sources of SRP. The components of Sum Sources (mineralization, external load, and entrainment) are shown to the right. Error bars represent one standard error around the mean, (n = 14 weeks). Input of SRP from precipitation and epilimnetic sediment release were < 1 mg SRP m"2 d"1 and are not shown here.

Page 109: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

98

CHAPTER 3

Estimates of phosphorus entrainment in Lake Mendota: A comparison of one-

dimensional and three-dimensional approaches

by

Amy M. Kamarainen1, Hengliang Yuan23, Chin-Hsien Wu2, and Stephen R.

Carpenter1

1 Center for Limnology, University of Wisconsin, 680 N. Park St., Madison,

Wisconsin, 53706 USA 2 Department of Civil and Environmental Engineering, University of Wisconsin, 1269D

Engineering Hall, 1415 Engineering Drive, Madison, Wisconsin, 53706 USA 3 Deep Water Engineering, Technip USA Inc., 11700 Old Katy Road, Suite 150,

Houston, Texas, 77079 USA

Keywords: three-dimensional hydrodynamic model, entrainment, phosphorus, lake,

physical model

Status: Kamarainen, A. M., H. Yuan, C. H. Wu, and S. R. Carpenter. 2009. Estimates

of phosphorus entrainment in Lake Mendota: A comparison of one-dimensional and

three-dimensional approaches. Limnology and Oceanography Methods 7:553-567.

Page 110: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

Acknowledgments

We thank David Armstrong and Nicholas Preston for helpful comments on the

draft manuscript. Also, we are indebted to Reinette Biggs, Daniel Collins, Owen

Langman, and Matt Van de Bogert for useful comments during the development of

the project. We owe a special thank you to Rachel Penczykowski for assistance with

field work and laboratory analysis. Also, James Thoyre, Ted Bier and Mark Lochner

provided help with field and laboratory work. Financial support for this project was

provided by NSF's North Temperate Lakes LTER program, a NSF fellowship to AMK,

and the A.W. Mellon Foundation.

Page 111: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

100 Abstract

Entrainment of phosphorus across the thermocline can be an important

nutrient source for phytoplankton in stratified lakes. In eutrophic stratified lakes,

seasonal entrainment can be responsible for delayed recovery following a decrease

in external phosphorus load. We compared seasonal estimates of entrainment

derived from single- and multi-location thermocline migration approaches.

Entrainment estimates from these methods were similar. A sampling approach based

on a single centralized location produced whole-lake entrainment estimates in line

with the mean value from the multi-location approach (the single-location approach

reasonably reproduced patterns at a weekly and seasonal time scale). In this study, it

was not essential to account for spatial variation in estimating annual rates of

entrainment. We also estimated entrainment based on a three-dimensional (3-D)

hydrodynamic model over a short period of significant thermocline migration. The 3-D

model was most useful in exposing spatial and temporal variation in temperature and

phosphorus profiles that are otherwise difficult to observe. Spatial variation in

phosphorus profiles was associated with upwelling of metalimnetic water represented

by the 3-D model. These transient dynamics, though a relatively small portion of an

annual phosphorus budget, may supply nutrients to epilimnetic phytoplankton during

periods of nutrient limitation. There is a need for further research that combines 3-D

hydrodynamic modeling with field collection of biological and chemical data.

Page 112: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

101 Introduction

Entrainment, herein defined as the turbulent exchange of water and solutes

across the lower boundary of the epilimnion, is a significant component of

biogeochemical fluxes to the epilimnion in many stratified lakes. Research on

entrainment has often focused on phosphorus (P) because of its role as a limiting

nutrient for phytoplankton and its importance in eutrophication. The relative

contribution of entrainment towards the total P supply to the epilimnion can vary

considerably from year to year (Effler 1986; Stauffer 1987; Soranno 1997).

Entrainment is one of the primary mechanisms by which P is recycled within the

water column of stratified lakes and explains some of the inter-annual variation in the

occurrence of algal blooms (Lathrop et al. 1999). Also, in some systems, entrainment

provides a substantial source of phosphorus to phytoplankton during the growing

season (Larsen et al. 1981; Kortmann et al. 1982; Hejzlar et al. 1993) and

entrainment at the end of the summer has been identified as a mechanism for

delayed recovery following restoration efforts aimed at decreasing external loads of P

(Effler et al. 1986; Jeppesen et al. 2005; Sondergaard et al. 2007).

Estimates of entrainment have often relied on the simplifying assumption that

temperature and P concentrations are homogeneous along a horizontal plane

(Soranno et al. 1997; Franke et al. 1999; Baehr and DeGrandpre 2004). This

assumption allows the monitoring of variables at a single location and substantially

simplifies the sampling and analysis. Previous publications address the limitations of

(Stauffer 1985; Stauffer 1993) and justifications for (Patterson et al. 1984) this

Page 113: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

102 assumption. A number of recent studies, however, show spatial heterogeneity in

turbulent events (Etemad-Shahidi and Imberger 2001; Saggio and Imberger 2001;

Maclntyre et al. 2002; Boegman et al 2003; Maclntyre et al. 2006; Na and Park

2006); and yet others connect dynamic physical processes to spatial heterogeneity in

biological and chemical variables (Ivey and Boyce 1982; Stauffer 1985; Maclntyre

and Melack 1995; Maclntyre and Jellison 2001; Eckert et al. 2002; Chao et al. 2006;

Marce et al. 2007).

While spatial heterogeneity has been demonstrated, transience is often used

as a rationale for averaging over spatial variation in field studies. Researchers

interested in seasonal averages and annual P budgets may justifiably average over

spatial heterogeneity at much shorter time scales. If one is interested in transient

phenomena such as algal blooms, however, accounting for temporally and spatially

heterogeneous nutrient fluxes likely becomes important. Eckert et al. (2002) revealed

changes in stratification patterns over a six-hour period and found the changes

differed among locations within the lake. These changes in physical properties were

reflected in changes in the dissolved oxygen, redox intensity, and hydrogen sulfide

profiles among locations.

Spatial and temporal variability in physical processes could have important

effects on estimates of P entrainment, and new modeling approaches may better

capture this variability compared to customary methods for estimating entrainment.

To evaluate this possibility, we used a spatially explicit sampling design to compare

single-location and multi-location estimates of seasonal P entrainment. We also

Page 114: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

103 applied a three-dimensional (3-D) hydrodynamic model to a six-day period of high

entrainment rates to assess the utility of this model in quantifying entrainment.

Materials and Procedures

Field and Laboratory Methods

Throughout the stratified portion of the summer of 2005 (23 June - 29

September), temperature, total phosphorus (P) profiles were collected over multiple

locations in the lake (Figure 1). Samples were taken at approximately two-week

intervals, and were collected more frequently following periods of strong wind.

Samples were consistently collected at the central station (in the deepest area of the

lake) throughout the summer, while an additional four sampling locations were

chosen in accordance with different wind patterns. All locations that were sampled at

least once are depicted in Figure 1.

Manual temperature profiles were collected at 0.5 -1 .0m intervals using a

YSI Dissolved Oxygen and Temperature probe (instrument model 58, probe model

5739, YSI Incorporated). Whole-water samples were collected in acid-washed

polyethylene bottles at 1 m - 4 m intervals using a peristaltic pump for determination

of total P concentrations, and stored on ice. P samples were preserved using Optima

HCI, while SRP samples were refrigerated and analyzed within 24 hours.

P samples were analyzed using a Technicon Auto Analyzer following

persulfate digestion (American Public Health Association et al. 1995). Complete

profiles at 1 m resolution were constructed for each sampling date and location using

Page 115: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

104 linear interpolation to estimate P concentrations for un-sampled depths.

Interpolated profiles were used in all analyses.

Conventional thermocline movement analysis

Total P transport across the lower boundary of the epilimnion may be

attributed to two processes 1) bulk P entrainment occurs episodically when P is

transported during thermocline migration and 2) P flux occurs continuously due to

turbulent diffusion across the epilimnetic boundary (molecular diffusion is considered

negligible). Bulk entrainment was quantified as the mass of P transported from the

metalimnion to the epilimnion when the thermocline deepened, computed as the

product of estimated volume of water entrained and the measured concentration of P

in entrained water prior to entrainment. The thermocline served as an estimate of the

boundary between the epilimnion and deeper waters. Therefore, we used the position

of the thermocline before and after entrainment to estimate the volume of water

entrained. We used manually collected temperature profiles to compute entrainment

so that our analysis would represent an approach customarily applied in both

research and management contexts (Effler et al. 1986; Stauffer 1987; Stauffer 1993;

Sorannoetal. 1997).

The thermocline was defined as the depth in the profile at which the maximum

temperature change occurs (Hutchinson and Edmondson 1957). A strong

temperature gradient, or diurnal thermocline, within five meters of the lake surface

(due to microstratification) was not considered in defining the depth of the

Page 116: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

105 thermocline. When the thermocline depth changed, the volume of water

transported was estimated from bathymetric data for the lake. This volume was

multiplied by the P concentration at depth at each sampling location to obtain the bulk

mass of P transported through entrainment at each location.

Additional transport of P across the thermocline boundary may occur due to

turbulent diffusion. This flux was quantified based on the following equations, which

represent an application of Fick's law where turbulent diffusive flux is directly

proportional to the concentration gradient and the eddy diffusion coefficient (Kz)

(Powell and Jassby 1974; Stauffer 1992, Stauffer 1993):

Jpz = -KzC'zAzAt

And

Kz = (1A/T^) - T^))) x (At^r1

j=z+l

Where: Jpz = the mean flux of phosphorus (p) in the z direction Kz = eddy diffusion coefficient (cm2 s"1) z = indicates the depth of the thermocline A = Lake area at surface (o), or at depth 1 m above or below the thermocline (j) T = temperature (°C) t = time step, period between sampling events

A single eddy diffusion coefficient and P gradient were defined at each location

based on a 2 m temperature gradient centered at the depth of the thermocline. The

mass of P transported through turbulent diffusion was estimated for the entire period

between sampling events, and this cumulative mass was added to the bulk

Page 117: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

106 entrainment estimated from thermocline migration to arrive at an estimate of the

total P mass transported across the thermocline boundary at each location. Total

entrainment was estimated at each location as though the location may be

considered representative of the entire lake (i.e. estimates of bulk volume transported

were based on total lake volume at depth). Estimates at each location were then

averaged to derive a lake-wide mean entrainment rate over the period between

sampling events.

We evaluated entrainment rate estimates by comparing these to the change in

epilimnetic TP mass. During time intervals when sedimentation is constant and net

inputs are small, changes in epilimnetic TP mass should be positively correlated with

entrainment rates. As presented below, these conditions were satisfied in Lake

Mendota over the summer of 2005. Mean epilimnetic TP concentrations were

calculated at all locations, multiplied by the volume of the epilimnion over the whole

lake, and presented as lake-wide arithmetic mean TP masses. All locations were

weighted equally in calculating the lake-wide mean value for entrainment and

epilimnetic TP mass.

Estimates of wet and dry deposition of P on the lake surface are minimal

during summer (mean = 8.7 kg d"1, range = 6.6 - 43.8 kg d"1). Estimates are based

on observed precipitation and estimated P concentration in precipitation of 0.032 mg

L"1 measured in the nearby Lake Wingra watershed (Kleusener 1972), along with an

estimated dry deposition rate of 0.62 kg TP ha"1 yr"1 from (Amy et al. 1974).

Page 118: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

107 Over 90% of external loading from the landscape occurred between

January and June during 2005, before the time interval analyzed here (Lathrop

2007). The external loading of P over the summer was on the order of 2,000 kg TP

(20.4 kg d"1), based on measured inputs scaled to the whole watershed according to

methods used in Lathrop et al. 1998. Meanwhile, monitored outflow data combined

with the mean P concentration of the epilimnion show that loss through the outflow is

approximately 15.8 kg d"1. Thus, net retention of P was approximately 4.6 kg d"1

during the summer of 2005.

Release of P from littoral sediments is another potential source of P that is

small in comparison to variation in estimates of entrainment and changes in

epilimnetic P mass. Previous studies show a mean epilimnetic P release rate for

Lake Mendota ranging from 1.2 to 3.5 mg P m"2 d"1 (Stauffer 1987). The total

contribution of littoral sediments should be between 17.4 and 51 kg P d"\ as

approximately 37% of the total sediment surface area may be considered littoral

(based on a mean thermocline depth of 9.5 m).

The total net input of P for 2005, considering aerial input, fluvial input and

output, and release from littoral sediments was between 30.7 and 64.3 kg P d"1. The

mean entrainment rate over the whole summer was 473 (± 281.2) kg P d"1. Thus, the

other sources of P are about 10% of the mean daily entrainment rate we measured.

While the magnitude of sedimentation is great in Lake Mendota, sedimentation

trap studies show that this flux is roughly constant over the summer period between

July and September (Fallon and Brock 1980). If sedimentation is constant it would

Page 119: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

108 have no effect on the correlation of entrainment with changes in epilimnetic P

mass. We conclude that the correlation coefficient of entrainment and total epilimnetic

P mass should be large and positive for an entrainment estimator that is accurate.

3-D hydrodynamic model

We accounted for the full suite of dynamic processes using a 3-D

hydrodynamic model for six days over the period from the 23rd (day 266) to the 29th

of September 2005 (day 272). The model could not be applied over a longer time

period due to computational costs, and this period was chosen due to high rates of

entrainment and observed spatial heterogeneity. The estimates of entrainment for

this six-day period were then compared to the single- and multi-location conventional

entrainment estimates derived for the same period.

The hydrodynamic model developed by Yuan and Wu (2004) employs the full

Reynolds averaged Navier-Stokes equations, free of the hydrostatic pressure

assumption that is commonly employed in other 3-D lake models (Hodges et al.

2000; Rueda and Schladow 2003). This allows for more accurate simulation of

internal wave evolution and dynamics that are important to solute transport and

energy transfer in the lake. The partial differential equations that govern mass and

momentum conservation are solved by a generalized implicit method, by which all

flow field components (such as three velocity components, pressure, free-surface

elevation, and temperature) are solved simultaneously at each time step. The model

employs a generic length scale approach for turbulence closure that explicitly

Page 120: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

109 includes four popular turbulence schemes, offering the advantage of selectable

parameterization to achieve an optimum result (Umlauf and Burchard 2003). Several

flux-limiting schemes in combination with direction splitting techniques are used to

calculate advection terms in scalar transport.

Following standard procedures (Dee 1995; Roache 1998), the model has been

carefully verified and validated with several free-surface flow problems (Yuan and

Wu, 2004; Yuan and Wu 2006; Yuan 2007). Also, we compared daily temperature

profiles produced by the model to empirical temperature profiles collected by

temperature loggers. Empirical temperature profiles were collected at one-minute

intervals using HOBO Underwater Temperature Data Loggers (0.02 °C resolution,

±0.2 °C accuracy, MicroDAQ.com, Ltd.) at the same locations as P sampling. The

data loggers were placed at 1 m depth intervals through the metalimnion and at 2m

intervals through the upper epilimnion and hypolimnion. Therefore, five thermistor

chains were positioned at different locations during the 3-D model focal period from

the 23rd (day 266) to the 29th of September 2005 (day 272). Our intent is to present

an assessment of the 3-D model as a tool in studying the physics of P recycling in a

natural lake, rather than demonstrate validation of the model. Therefore, further

technical details of the model and its verification and validation can be found in Yuan

and Wu 2004, Yuan and Wu 2006, and Yuan 2007.

In model simulations, the horizontal grid size was 100 m; grid-refinement tests

showed that such a resolution gives grid-independent results. At this resolution there

are approximately 7,200 horizontal "sampling locations" within the lake. A subset of

Page 121: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

110 the 7,200 locations was used, however, because at depths shallower than 12m

the water column may not have been stratified and the thermocline could not be

defined. In model simulations, 15 vertical layers were used and a time step of 100

seconds was chosen. Such a set of numerical parameters lead to a ratio of

simulation time to real time of approximately 1:10.

The model was driven by meteorological data, including: wind speed and

direction, air temperature, solar radiation, barometric pressure, and relative humidity

(Figure 2). Those data were obtained from the database of the Rooftop Instrument

Group measured on the Atmospheric, Oceanic and Space Science Building at

University of Wisconsin - Madison (approximately 0.5 km south of Lake Mendota,

http://rig.ssec.wisc.edu/).

The model was initialized at 0000 hours on day 265 (allowing for a 24-hour

burn in period before use of model output), with a null velocity field and a temperature

profile obtained by averaging the five empirical temperature profiles. Coefficients for

the surface and bottom drag were selected according to Wuest and Lorke (2003). A

zero flux boundary condition was specified at the lake bottom for temperature

calculation. The attenuation coefficient was specified according to the Secchi depth

(approximated 3 m, measured during the study period), following the method of

Beeton(1958).

Assessment

Study Site

Page 122: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

111

Lake Mendota is a stratified, dimictic lake in southern Wisconsin with a

surface area of 39.1 km2, mean depth of 12 m and maximum depth of 24 m. P inputs

are primarily associated with runoff from the agriculturally-dominated watershed

(Lathrop et al. 1998). The lake is stratified from the middle of June until the middle of

October. P builds up in the metalimnion and hypolimnion of the lake each summer,

while SRP concentrations in the epilimnion often fall below detection limits.

General Outcome of Conventional Analysis

Entrainment over the stratified period between late-June (day 174) and mid-

September (day 266) resulted in the total gross transport of approximately 5,296 kg

TP. There was considerable spatial heterogeneity in the estimates of entrainment,

with the greatest spatial heterogeneity following periods of windy weather (day 206-

207, day 266-272) (Figure 3). Entrainment was negative on a number of occasions;

this indicates that the lower boundary of the epilimnion became shallower due to still

wind conditions and high solar insolation. The magnitude of entrainment was

positively correlated with mean wind speed (r=0.79, p=0.012, Pearson's Correlation)

and negatively correlated with the mean solar flux over the period between sampling

events (r=-0.93, p<0.001, Pearson's Correlation) (Figure 4).

The majority of entrainment occurred late in the season during a period of

persistent winds. Over six days in late September (day 266-272) approximately

10,204 kg TP was entrained into the epilimnion, resulting in a mean daily rate of

1,701 kg TP day"1 (Figure 3). As the thermocline degraded prior to fall mixis, a further

Page 123: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

112 12,482 kg TP was incorporated into the epilimnion between 29 September 2005

(day 272) and 15 October 2005 (day 288). In total, conventional analyses indicate

that entrapment accounted for the vertical flux of approximately 27,982 kg TP during

the summer season (day 174 - day 288). Because spatially explicit sampling did not

continue after 29 September 2005, entrainment through the October period was not

considered in the remainder of the assessment.

Despite significant spatial heterogeneity in estimates of entrainment, the

samples collected at the central location generally coincided with the mean estimate

of entrainment for each sampling event. The entrainment observed at the central

station tended to be less extreme than estimates at other locations (Figure 3). When

summed over the entire summer (174 - 272), the total estimate of entrainment based

on the central location (15,653 kg P) was close to the multi-location estimate (15,500

kg P). These results provide support for the idea that sampling at a single central

location provides an acceptable estimate of the mean annual entrainment rate.

The validity of conventional entrainment estimates was supported by

comparison with estimates of the change in epilimnetic TP mass (Figure 5). We did in

fact find a significant positive correlation between estimated entrainment and the

observed change in the epilimnetic TP mass (r = 0.6, p < 0.001, Pearson's

Correlation). Note that estimates of entrainment tended to be greater than observed

changes in epilimnetic TP mass (i.e. most entrainment values fall above the 1:1 line).

This indicates that sedimentation is in fact a significant flux, and the amount lost from

Page 124: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

113 the epilimnion through sedimentation was, at times, greater than that supplied

through entrainment.

Sensitivity Analysis

Several definitions have been used to characterize the boundary between the

mixed layer and deeper waters, and we explored the effect of boundary definition on

the estimate of entrainment derived from the conventional approach. The boundary is

most often defined as the thermocline, or the depth at which the maximum

temperature change occurs (Max_Change) (Hutchinson and Edmondson 1957). Yet,

this approach may lead to spurious changes in thermocline and epilimnetic depth

(Soranno et al. 1997; Fee et al. 1996). Therefore we used three additional

approaches to define the epilimnetic boundary, including the depth of the primary

isotherm (Isotherm). The isotherm is estimated as the mean seasonal temperature at

the thermocline when the thermocline is defined by the Max_Change method

(Soranno et al. 1997; Robertson 1989, Marce et al. 2007). This approach tends to

limit the spurious changes in thermocline depth that may be observed with other

approaches (Soranno et al. 1997). In our study the primary isotherm was 18.3 °C for

the summer season. The second approach used was the Mid-point method, in which

the thermocline depth is halfway between the base of the epilimnion and the top of

the hypolimnion (Mid_Point) (Soranno 1995). Here, the base of the epilimnion and

top of the hypolimnion are both defined as the first depth at which the temperature

change is greater than one degree Celsius per meter, starting at the lake surface or

Page 125: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

114 lake bottom respectively. Finally, we also used a more conservative measure of P

flux based on changes in the maximum depth of the mixed layer (Within_One). In this

case, the mixed layer was defined as the mass of water with a uniform temperature

(i.e. temperature differential within one degree Celsius) (Nagai et al. 2005).

Estimates of entrainment differed slightly depending on the approach used for

defining the boundary of the mixed layer. Mean rates and the range of estimates

were similar in all cases (Figure 6). Further, the different approaches resulted in

entrainment estimates that were positively correlated with one another (r > 0.72, p <

0.01, in all cases). Within One was typically smaller in magnitude than the others,

which would be expected because it is the most conservative definition. Yet, the Max

Change thermocline definition matched best with observed changes in the epilimnetic

P mass, therefore, the Max_Change definition was used to summarize conventional

estimates of entrainment over the stratified season.

We also examined whether spatial variation in entrainment observed during

the 206-207 and 266-272 periods may be exclusively local and therefore lead to

overestimation of lake-wide entrainment when scaled to the whole lake. To do so, we

used Thiessen polygons (delineated using ArcGIS) to define the area of influence

around each sampling point. Entrainment estimates were then weighted by the

relative contribution of each site to the total lake estimate (weighting = surface area of

Thiessen polygon/total surface area of lake). We found no difference between

estimates derived on a whole-lake basis compared to those derived using the

Thiessen polygon weighting.

Page 126: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

115

Hydrodynamic Modeling Approach

We used the model to estimate the volume of water associated with

thermocline migration during the period from 23 September (day 266) until 29

September (day 272). During this period the thermocline migrated over 2 m to a

lake-wide mean depth of 13 m. In particular, a storm occurred on day 271 into 272.

During this storm, cool air temperatures and high wind speeds (mean ~ 10m/s) from

the west and northwest contributed to thermocline deepening (Figure 7).

The hydrodynamic model reasonably reproduced thermal profiles for the

period of interest (Figure 8). The model results do show slightly warmer temperatures

in the hypolimnion than those observed in the temperature profile. This may be

because we used a zero flux boundary condition in the model. Our assumption that

heat flux at the lake bottom was zero over this six-day period may be erroneous, yet

the hypolimnetic warming in the model did not appear to affect thermocline definition

(defined as the depth of max rate of temperature change (Max Change)). The six-day

range in the depth of the thermocline defined by the 3-D model (11.1 - 13.6 m)

conformed reasonably well with the beginning (day 266) and ending (day 272)

thermocline depths defined by the conventional method (10.8-13 m). Thus, we

determined that the thermal structure of interest was indeed captured by the 3-D

model, and the thermal structure defined by the model was used to estimate

entrainment over the six-day period.

Page 127: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

116 In order to compare the 3-D model output with the conventional approach

(both single- and multi-location approaches), all approaches were used to estimate

entrainment from 23 September 2005 until 29 September 2005 (day 266 - day 272).

Estimates for this period using the single-location conventional method resulted in an

estimate of 10,836 kg TP, the multi-location conventional method resulted in an

estimate of 10,204 (± 3340) kg TP, while the 3-D model provided an estimate of

12,550 kg (± 1176) TP (Figure 9). The estimates of entrainment based on

conventional and 3-D model calculations resulted in slight, though not statistically

significant (t-test, p > 0.4, d.f. = 9), overestimation of the change in TP mass that was

observed in the epilimnion (7,806 (± 1087) kg TP) over the 23 September - 29

September period (single-location results not included in statistical test due to small

sample size (n=1)) (Figure 9).

The 3-D model captures upwelling dynamics likely responsible for spatial

variation in P entrainment observed on dates associated with storm conditions.

Representation of the lake thermal structure in the model demonstrates that under

high wind speeds (~10 m s"1) spatial variation in water temperatures is directly related

to spatial variation in P concentrations across locations in the lake (Figure 10). The

constant westerly winds blowing from days 270 through 273 resulted in upwelling of

metalimnetic water in the western basin of the lake. Spatial heterogeneity in TP and

SRP concentrations at depth, likely caused by upwelling, is apparent along the

boundary between epilimnetic and metalimnetic water (11 - 13 m) (Figure 10).

Similarly, a horizontal cross-section across the 12 m depth plane (Figure 11 b)

Page 128: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

117 demonstrates that the thermal structure of the lake responds to wind speed and

direction, and this spatial variation in temperature is reflected in P concentrations at

specific locations (Figure 11). Such spatial variation in temperature and P profiles

may give rise to variation in estimates of entrainment. Congruence between thermal

structure represented in the model and observed spatial variation in P profiles

demonstrates that the model is capturing dynamics relevant to the distribution of P

over short time scales.

While gross entrainment rates derived from all three approaches did not differ

significantly during this six-day period, we were interested in whether the three

different approaches converged on similar estimates of entrainment at a daily time

scale. We used the P profile on day 266 along with daily temperature profiles from

the temperature loggers to quantify daily P flux as the thermocline eroded. We found

there were marked differences in daily entrainment rates among the three

approaches, and daily entrainment rates were not correlated with one another. Given

that we sampled P profiles less frequently than temperature profiles, we were not

able to assess which of these approaches most closely approximated changes in

epilimnetic P mass at a daily time step.

Discussion

Studies of P entrainment have shown that 1.) entrainment may represent a

significant contribution to the annual P budget of the photic zone in many lakes and

2.) entrainment may provide a source of P to P-limited planktonic organisms during

Page 129: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

118 the summer growing season (Sondergaard et al. 2003). Our study suggests that

single and multi-location approaches deliver comparable estimates of seasonal

entrainment. The 3-D model represented thermal structure and hydrodynamic

patterns relevant to P distribution and transport over a short period of thermocline

deepening. Also, 3-D model estimates of gross entrainment over a six-day period

showed general agreement with the conventional estimates, though total entrainment

quantified using the 3-D was slightly higher. Estimates of daily entrainment differed

among conventional and 3-D approaches and were not correlated. The incongruence

among daily estimates based on three approaches suggests validating a P

entrainment model at a daily time scale requires more intensive measurements.

Our results corroborate earlier findings that suggest horizontal variation in P

concentrations is a relatively small source of overall variation in estimates of P stored

in the metalimnion (Stauffer 1985). Our seasonal estimate of entrainment captured at

a single central location is consistent with estimates based on the multi-location

conventional sampling method. These results generally hold true even when the

definition of the thermocline is modified. In this study, sampling locations were

distributed on either side of the node of the internal seiche such that we typically

captured both inflated and deflated P concentrations during thermocline tilting. The

effective averaging of these values canceled out the effects of thermocline

displacement on the estimates of entrainment for a single date. Because estimates of

entrainment at the central station correspond with spatially explicit estimates, and the

central station is less prone to the confounding effects of thermocline tilt, we suggest

Page 130: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

119 that sampling at a single location is effective in characterizing lake-wide amounts

of P entrainment into the epilimnion. Thus, previous efforts to estimate mean annual

entrainment based on a single centralized sampling location are likely justified,

particularly in lakes with little morphometric variation.

Entrainment predominantly occurs in large pulses concurrent with atmospheric

cooling and the passage of storms (Stauffer and Lee 1973; Stauffer 1993; Imberger

and Patterson 1990; Soranno et al. 1997). Consistent with these findings the largest

entrainment rates in our study occurred during periods of windy weather (Figure 4).

Our sampling approach (regular sampling combined with increased sampling

frequency following high wind speeds) was designed to capture the majority of

entrainment events. The general thermal trend over the summer suggests that big

changes in the depth of the thermocline were not frequent during the summer of 2005

(Figure 2). Instead thermocline migration could be characterized as gradual, and we

believe our sampling routine was sufficient to capture entrainment associated with

this gradual progression of the thermocline.

At the same time, our seasonal analysis demonstrates that entrainment is not

limited to large storm events but that a significant flux also occurs during relatively

calm periods due in large part to turbulent diffusion across the thermocline (Figure

12). It is difficult to account for this flux because Kz values can be highly variable and

tend to increase when entrainment intensifies. Thus, it may be difficult to capture

short-term changes in Kz using a weekly monitoring approach, but such short-term

changes in Kz, and the associated P flux, can be accounted for by the 3-D model.

Page 131: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

120 Therefore, though mean seasonal estimates of entrainment can be accomplished

using the single-location conventional approach, many of the physical dynamics

responsible for variation in entrainment estimates are better represented using a 3-D

modeling technique.

The three-dimensional approach is most useful in understanding the

connection between physical processes and the lateral distribution of P. 3-D

hydrodynamic modeling lends insight into timing of delivery of P over short time

scales and can identify local regions of upwelling responsible for transient spatial

variation in P concentrations. During periods of upwelling, transient supplies of P can

become available at specific locations within the lake (Western basin of Lake

Mendota, day 272 - day 273, Figure 8). Due to high uptake rates for SRP and high

rates of lateral movement of solutes within the epilimnion it is extremely difficult to

observe local, transient changes in P concentrations. Yet, these discrete fluxes of P

may supply P-limited phytoplankton with needed nutrients. The 3-D hydrodynamic

model effectively captures the physical dynamics likely responsible for transient P

fluxes.

While we observed spatial heterogeneity in P profiles during the period

between 266-272, one limitation of this sampling design was that we were not able to

more closely examine the contribution of fluxes near the lateral boundaries of the

lake. All of our sampling locations were positioned in water that was at least 12 m

deep, and thus any contribution of P to the epilimnion through mixing at the lateral

boundaries could only be recognized following advection of dissolved and suspended

Page 132: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

121 material to our sampling locations. Recent work suggests boundary mixing is

important to scalar transport in lakes ((Etemad-Shahidi and Imberger 2001,

Maclntyre et al. 2002). Our 3D hydrodynamic model is capable of simulating internal

wave dynamics (Yuan 2007), and provides a platform to study turbulent mixing in

benthic boundary layer and its interaction with internal waves. With a sediment

transport subroutine added, the present 3D hydrodynamic model can be used to

further study P transport due to lateral boundary mixing.

Our finding that a single-location thermocline migration approach is sufficient

to capture the seasonal trend in entrainment will be useful in a management context.

Meanwhile, our observations of the relationship between physical processes and

spatial heterogeneity in P profiles will be of interest to researchers who study

transient dynamics associated with bloom formation and phytoplankton patchiness.

While it may be possible to use a single-location conventional approach to achieve a

reasonable accounting for the total entrainment at a seasonal scale, a more complex

model is required if one wishes to understand the mechanisms responsible for the

transport of P. The non-hydrostatic nature of our 3D hydrodynamic model allows for

more accurate simulation of a variety of physical processes such as flow over steep

topography, upwelling, deep water convection, dynamics of short surface waves,

degeneration of basin-scale waves, and evolution of nonlinear internal waves. All of

theses processes affect (to differing degrees) P transport in lakes. Thus, the 3-D

model is a useful tool for representing physical processes that generate mixing and

transport of biologically important solutes over intra-seasonal time scales.

Page 133: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

122

Comments and Recommendations

While our results suggest that sampling at a single central location is sufficient

for capturing a seasonal estimate for entrainment, readers should keep in mind the

morphometric complexity of the lake when applying this conventional method to other

systems (Robarts et al. 1998). Lake Mendota has a relatively simple basin shape.

Yet, studies have shown that the shallow sill, which forms some separation between

the main lake basin and the western arm of the lake, is sufficient to impose some

bias in solute concentrations between the two basins (Stauffer 1985). Therefore,

morphometric complexity could lead to greater spatial variation in entrainment

estimates. The 3-D model could be useful in exploring the spatial variation in

entrainment estimates in lake basins with greater morphometric complexity.

Time-dependent, three-dimensional computer models are capable of resolving

the high degree of spatial and temporal variability inherent to lake dynamics, such as

nutrient entrainment. Due to computational costs, however, the application of 3-D

models to real lakes is restricted to coarse grid simulations, and the associated

numerical issues must be carefully addressed to ensure the quality of model results.

Further, even the coarse-grid simulation is fairly time intensive, and sometimes

becomes impractical for the long-term (e.g. seasonal or annual time scale)

application in large lakes. Therefore, we recommend a combined approach, including

field measurements, conventional limnological methods, and 3-D modeling to

investigate entrainment over multiple spatial and temporal scales. At the same time,

Page 134: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

123 we recognize opportunities to improve the efficiency of the 3-D model (e.g.

through code parallelization), and to broaden its applicability through explicit coupling

with chemical and biological models.

In conclusion, the 3-D hydrodynamic model agrees with conventional

estimates of entrainment and both generally agree with observed changes in

epilimnetic P mass over a short period of late-summer thermocline migration. The

simplest approach involving sampling at a single central location, near the deepest

part of the lake, resulted in sufficient estimates of mean annual entrainment. The

hydrodynamic model, however, can provide fine-scale information about the spatial

and temporal variation in P concentrations that is not achievable in other ways. In this

regard, further research and development to expand the use and applicability of 3-D

hydrodynamic models would be productive.

Page 135: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

References 124

American Public Health Association., American Water Works Association., and Water Pollution Control Federation. 1995. Standard methods for the examination of water and wastewater, 19th ed. American Public Health Association.

Amy, G. R. Pitt, Rameshawar-Singh, W. L. Bradford and M. B. LaGraff. 1974. Water quality management planning for urban runoff. EPA 440/9-75-004. U.S. Environmental Protection Agency, Washington, D.C.

Baehr, M. M., and M. D. Degrandpre. 2004. In situ pCO(2) and 0-2 measurements in a lake during turnover and stratification: Observations and modeling. Limnol. Oceanogr. 49: 330-340.

Beeton AM. 1958. Relationship between secchi disk readings and light penetration in Lake Huron. Trans. Am. Fish. Soc. 87: 73-79.

Boegman, L, J. Imberger, G. N. Ivey, and J. P. Antenucci. 2003. High-frequency internal waves in large stratified lakes. Limnol. Oceanogr. 48: 895-919.

Chao, X. B., Y. F. Jia, C. M. Cooper, F. D. Shields, and S. S. Y. Wang. 2006. Development and application of a phosphorus model for a shallow oxbow lake. J. Environ. Eng. -Asce 132: 1498-1507.

Dee, D. P. 1995. A pragmatic approach to model validation, p. 1-13. In D. R. Lynch and A.M. Davies [eds.], Quantitative skill assessment for coastal ocean models.

Eckert, W., J. Imberger, and A. Saggio. 2002. Biogeochemical response to physical forcing in the water column of a warm monomictic lake. Biogeochemistry 61: 291-307.

Effler, S. W., M. C. Wodka, C. T. Driscoll, C. Brooks, M. Perkins, and E. M. Owens. 1986. Entrainment-Based Flux of Phosphorus in Onondaga Lake. J. Environ. Eng.-Asce 112: 617-622.

Etemad-Shahidi, A., and J. Imberger. 2001. Anatomy of turbulence in thermally stratified lakes. Limnology and Oceanography 46:1158-1170.

Fallon, R. D., and T. D. Brock. 1980. Planktonic Blue-Green-Algae - Production, Sedimentation, and Decomposition in Lake Mendota, Wisconsin. Limnology and Oceanography 25: 72-88.

Page 136: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

125 Fee, E. J., R. E. Hecky, S. E. M. Kasian, and D. R. Gruikshank. 1996. Effects of

lake size, water clarity, and climatic variability on mixing depths in Canadian Shield lakes. Limnol. Oceanogr. 41: 912-920.

Franke, U., K. Hutter, and K. Johnk. 1999. A physical-biological coupled model for algal dynamics in lakes. Bull. Math. Biol. 61: 239-272.

Hejzlar, J., M. Balejova, D. Kafkova, and M. Ruzicka. 1993. Importance of Epilimnion Phosphorus Loading and Wind-Induced Flow for Phytoplankton Growth in Rimov Reservoir. Water Sci Technol 28: 5-14.

Hodges, B. R., J. Imberger, A. Saggio, and K. B. Winters. 2000. Modeling basin-scale internal waves in a stratified lake. Limnol. Oceanogr. 45:1603-1620.

Hutchinson, G. E., and Y. H. Edmondson. 1957. A treatise on limnology. Wiley.

Ivey, G. N., and F. M. Boyce. 1982. Entrainment by Bottom Currents in Lake Erie. Limnol. Oceanogr. 27:1029-1038.

Imberger, J., and J. C. Patterson. 1990. Physical Limnology. Adv. App. Mech. 27: 303-475.

Jeppesen, E., M. Sondergaard, J.P. Jensen, K.E. Havens, O. Anneville, L. Carvalho, M.F. Coveney, R. Deneke, M. Dokulil, B. Foy, D. Gerdaux, S.E. Hampton, S. Hilt, K. Kangur, J. Kohler, E.H.H.R. Lammens, T.L. Lauridsen, M. Manca, M.R. Miracle, B. Moss, P. Noges, G. Persson, G. Phillips, B. Portielje, S. Romo, C.L. Schelske, D. Straile, I. Tatrai, E. Willen, and M. Winder. 2005. Lake responses to reduced nutrient loading - an analysis of contemporary long-term data from 35 case studies. Freshwater Biol. 50: 1747-1771.

Kluesener, J. W. 1972. Nutrient transport and transformations in Lake Wingra, Wisconsin. Ph.D. Thesis. Water Chemistry Program, University of Wisconsin, Madison. 181 p.

Kortmann, R. W., D. D. Henry, A. Kuether, and S. Kaufman. 1982. Epilimnetic Nutrient Loading by Metalimnetic Erosion and Resultant Algal Responses in Lake Waramaug, Connecticut. Hydrobiologia 91: 501-510.

Larsen, D. P., D. W. Schults, and K. W. Malueg. 1981. Summer Internal Phosphorus Supplies in Shagawa Lake, Minnesota. Limnology and Oceanography 26: 740-753.

Page 137: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

126 Lathrop, R.C., S.R. Carpenter, C.A. Stow, P.A. Soranno and J.C. Panuska. 1998.

Phosphorus loading reductions needed to control blue-green algal blooms in Lake Mendota. Can. J. Fish. Aquat. Sci. 55: 1169-1178.

Lathrop, R. C, S. R. Carpenter, and D. M. Robertson. 1999. Summer water clarity responses to phosphorus, Daphnia grazing, and internal mixing in Lake Mendota. Limnol. Oceanogr. 44:137-146.

Lathrop, R. C. 2007. Perspectives on the eutrophication of the Yahara Lakes. Lake and Reservoir Management 23: 345-376.

Maclntyre, S., and J. M. Melack. 1995. Vertical and horizontal transport in lakes: Linking littoral, benthic, and pelagic habitats. J. N. Am. Benthol. Soc. 14: 599-615.

Maclntyre, S., and R. Jellison. 2001. Nutrient fluxes from upwelling and enhanced turbulence at the top of the pyncocline in Mono Lake, California. Hydrobiol. 466: 13-29.

Macintyre, S., J. R. Romero, and G. W. Kling. 2002. Spatial-temporal variability in surface layer deepening and lateral advection in an embayment of Lake Victoria, East Africa. Limnology and Oceanography 47: 656-671.

Maclntyre, S., J. O. Sickman, S. A. Goldthwait, and G. W. Kling. 2006. Physical pathways of nutrient supply in a small, ultraoligotrophic arctic lake during summer stratification. Limnol. Oceanogr. 51:1107-1124.

Marce, R., C. Feijoo, E. Navarro, J. Ordonez, J. Goma, and J. Armengol. 2007. Interaction between wind-induced seiches and convective cooling governs algal distribution in a canyon-shaped reservoir. Freshwater Biol 52:1336-1352.

Na, E. H., and S. S. Park. 2006. A hydrodynamic and water quality modeling study of spatial and temporal patterns of phytoplankton growth in a stratified lake with buoyant incoming flow. Ecol. Modell. 199: 298-314.

Nagai, T., H. Yamazaki, H. Nagashima, and L. H. Kantha. 2005. Field and numerical study of entrainment laws for surface mixed layer. Deep-Sea Res. 52:1109-1132.

Patterson, J. C, P. F. Hamblin, and J. Imberger. 1984. Classification and Dynamic Simulation of the Vertical Density Structure of Lakes. Limnol. Oceanogr. 29: 845-861.

Page 138: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

127

Powell, T., and A. Jassby. 1974. Estimation of Vertical Eddy Diffusivities Below Thermocline in Lakes. Water Resour Res 10:191-198.

Roache PJ. 1998. Verification of codes and calculations. AIAA Journal 36: 696-702.

Robarts, R.D., Waiser, M.J., Hadas, O., Zohary, T., and S. Maclntyre. 1998. Relaxation of phosphorus limitation due to typhoon-induced mixing in two morphologically distinct basins of Lake Biwa, Japan. Limnol. Oceanogr. 43(6):

Robertson, D. M. 1989. The use of lake water temperature and ice cover as climatic indicators. Ph.D. thesis. Univ. of Wisconsin, Madison.

Rueda, F. J., and S. G. Schladow. 2003. Dynamics of large polymictic lake. II Numerical simulations. J. Hydraul. Eng. -Asce 129: 92-101.

Saggio, A., and J. Imberger. 2001. Mixing and turbulent fluxes in the metallimnion of a stratified lake. Limnol. Oceanogr. 46: 392-409.

Sondergaard, M., Jensen, E., and Jeppesen, E. 2003. Role of sediment and internal loading of phosphorus in shallow lakes. Hydrobiol. 506:135-145.

Sondergaard, M., Jeppesen, E., Lauridsen, T.L., Skov, C, van Nes, E.H., Roijackers, R., Lammens, E., and Portielje, R. 2007. Lake restoration: successes, failures and long-term effects. J. of Appl. Ecol. [doi 10.1111/J.1365-2664.2007.01363.x]

Soranno, P. A. 1995. Phosphorus cycling in the Lake Mendota ecosystem: internal versus external nutrient supply. Ph.D. thesis. Univ. of Wisconsin, Madison.

Soranno, P. A., S. R. Carpenter, and R. C. Lathrop. 1997. Internal phosphorus loading in Lake Mendota: response to external loads and weather. Can. J. Fish. Aquat. Sci. 54: 1883-1893.

Stauffer, R. E. 1985. Lateral Solute Concentration Gradients in Stratified Eutrophic Lakes. Water Resour. Res. 21 : 554-562.

—. 1987. Vertical Nutrient Transport and Its Effects on Epilimnetic Phosphorus in 4 Calcareous Lakes. Hydrobiol. 154: 87-102.

—. 1992. Efficient Estimation of Temperature Distribution, Heat-Storage, Thermocline Migration and Vertical Eddy Conductivities in Stratified Lakes. Freshwater Biol 27: 307-326.

Page 139: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

128

—. 1993. Effects of an Early Fall Cold-Front on Heat, Phosphorus, Silica, and Manganese Distributions in the Hypolimnion of Lake Mendota, Wisconsin. J. Hydrol. 151: 1-18.

Stauffer, R.E. and G.F. Lee. 1974. The role of thermocline migration in regulating algal blooms. In Modeling the Eutrophication Process. Ann Arbor Science Publishers Inc., Ann Arbor, Ml, USA

Umlauf L and Burchard H. 2003. A generic length-scale equation for geophysical turbulence models. J. Mar. Sys. 61: 235-265.

Wuest, A., and A. Lorke. 2003. Small-scale hydrodynamics in lakes. Annu. Rev. Fluid Mech. 35:373-412.

Yuan, H. 2007. Hydrodynamics and Scalar Transport in Two Madison Lakes. Ph.D. thesis. Univ. of Wisconsin, Madison.

Yuan, H., and C. H. Wu. 2004. An implicit 3D fully non-hydrostatic model for free-surface flows. Int. J. Numer. Meth. Fluids 46: 709-733.

Yuan, H., C. H. Wu. 2006. Fully non-hydrostatic modeling of surface waves. J. Eng. Mech. 132:447-456.

Page 140: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

>

1 ^

I 1 km

i?

' & •

20 •c

^ "

4 8 '2 16 30

'entra Station so n>

'?4

Central

?4

i i » «

V V.

s /'-''

Figure 1. Bathymetric map of Lake Mendota. Bathymetric lines are labeled with depth in meters. The twelve sampling locations are designated by dots, with the central station location designated by a star. Coordinates: 43°06 N, 89°25 W

Page 141: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

130

-o 6 -0)

& 4 -C S 2V

; (M^MAI^^

18 20 22 Water temperature (Degree)

Figure 2. Season-long air temperature, wind speed and water temperature for Lake Mendota for the summer season 2005.

Page 142: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

131

5000

4000 -\

3000

2000

1000

0

-looo H

-2000

-#-T-f-

ofc

£

-i?

# Other Locations i^ Central Station

—~ Mean

& •

T - & •

+•

r -

\ u

• ^

s * o? A

\ .# \ * o9 to ^ '

tf> «rs ^ A^

^ £> <** >$

<o

Time period (Day of year)

Figure 3. Daily entrainment rate for total phosphorus (TP) based on one to five locations. Negative entrainment values indicate that the thermocline was shallower at the end of the time period.

Page 143: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

>> OS -o O)

_*: > * -< '

c 0 E c

"ro

Ent

r

o lO

1

o

250

1500

4.0 4.5 5.0 5.5 6.0

Wind Speed (ms~1)

150 200 Solar Flux (Wm"2d_1)

300

Figure 4. Correlation between entrainment rates and mean wind speed or mean solar flux over the period between sampling events, r = Pearson's product moment correlation coefficient, p = p-value based on a = 0.05.

Page 144: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

133

LU

O O s

o o

—' O

1 8

Q - t

8 o in

o o

-10000 -5000 0 5000 10000 15000

Change in Epiiimnetic P (kg)

Figure 5. Positive correlation between the observed change in epilimnetic phosphorus mass at each location and the estimate of the mass of phosphorus entrained (r = 0.6, p <0.001).

Page 145: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

134

o o o

T5 £JJ

-S£ *—r

X .3 c= £L

O o o

o o sr>

o ~\

o

O

WO MC MP IT

Mixed-layer boundary definition

Figure 6. A comparison of mean entrainment rates based on four definitions of the boundary of the mixed layer. WO = Within One, MC = Max Change, MP = Mid-Point, IT = Isotherm.

Page 146: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

135

'265 267 268 269 270 271 272 273

Dav of the Year

Figure 7. Meteorological conditions between 23 September 2005 (day 266) and 29 September 2005 (day 272). Wind direction is from the north at 0 degrees, east at 90 degrees, south at 180 and west at 270. Each color in the temperature graph represents a depth within the thermal profile. Air and surface water temperatures show a cooling trend. Entrainment can be observed in the temperature graph as the depths represented by orange, yellow, and green are incorporated into the epilimnion. Strong west winds on day 271 and 272 are accompanied by rapid deepening of the thermocline.

Page 147: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

136

-5

10

15

20

• 4

X • y /

•J

[ Day 267

10 15 20 25 10 15 20 25 10 15 20 25

Temperature (°C) 10 15 20 25

-6

| - 1 0

Q .

-20

1

• /»

• Day 270

10 15 20 25 10 15 20 25

Temperature ("C)

-5

10

15

20

i a

i \ i

• / * S

* / / Day 272

10 15 20 25

Figure 8. Model validation of temperature profiles. Dots represent measured temperatures. Lines represent model fit to the data.

Page 148: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

137

C3 1

C3

a

Epilimnion 3D Model MuitUoc Single foe.

Figure 9. Cumulative estimate of phosphorus flux during the period between 23 September and 29 September (Day 266 - 272). "Epilimnion" represents an estimate of the change in phosphorus mass in the epilimnion over the period of interest; 3D Model" = the sum of phosphorus entrained according to the 3D hydrodynamic model; "Multi-loc." = the sum of phosphorus entrained according to the multi-location conventional approach; "Single-loc." = the sum of phosphorus entrained according to the single-location conventional approach where the single location is positioned near the center of the lake. Error bars represent the standard error around the mean of all locations (n=5 in all cases except Single-loc (n=1)).

Page 149: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

138

Day 271

2 3

Da\ 272

• "- ;

Velocity (ITS %) 0.1

Temperature O'C) ~ 2 0

18 : i6 14

B 1 2

Das 273

Location

Distance (in)

Location 2 Location 3

0.0 02 0.4 0.6 0.8

TP Concentration (mg L"1)

Figure 10. Thermal representation of a west to east cross-section of Lake Mendota during a three day period in late September (28 September (day 271) - 30 September (day 273)) based on 3-D hydrodynamic model. The images depict the thermal structure at 0000 hours on each date. The colors represent temperature in degrees Celsius. The arrows represent direction and magnitude of water velocity. Upwelling (cool metalimnetic water reaches the surface) can be observed at the west end of the lake on day 272 and 273. Total phosphorus (TP) concentration profiles are depicted for the three sampling locations indicated by vertical lines on the day 272 color image. Horizontal lines indicate a subset of the depths at which phosphorus samples were collected. Spatial variation in phosphorus profiles is apparent, particularly near the depth of the thermocline (thermocline ~ 13 m on day 272).

Page 150: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

139 7500

500(1

2500

2500 5000 7500

7500

5000

2500^

Vclocilx (tn/s) i l l"

Temperature (°C)

• * 2 0 19 IS

: |7

16 -Jl5

2500 5000

x(m)

7500

Figure 11. Spatial variation in thermal structure at the surface (A) and at 12 m below surface (B). Colors represent temperature in degrees Celsius. Thermal images are based on a horizontal cross-section of Lake Mendota using a 3-D hydrodynamic model for a 24-hour period on the 29th September 2005 (day 272). Numbers in the graphics represent the SRP concentration (mg/L) at each depth and location on day 272.

Page 151: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

1

2500

2000 i

— 1500 r-

T5

cS 1000 X

°- 500 •

-500

^ M Turbulent Flux l l Bulk Entrainment

^ *& <$> T& tf> ,$> r$> q& jV < 1 N ' N<&&' N Q p ' ^ ^ V ^ V ^ 0 ' ,£,<&' a<o<6'

Time Period (Day of Year)

Figure 12. Relative contribution of turbulent flux and bulk entrainment toward the total estimate of entrainment for each sampling period.

Page 152: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

141

CHAPTER 4

Long-term trends in ice cover, stability, phosphorus and water quality in

eutrophic Lake Mendota

by

Amy M. Kamarainen1,3, Richard C. Lathrop1,2 and Stephen R. Carpenter1

1 Center for Limnology, University of Wisconsin, Madison, 680 N. Park St., Madison,

Wl 53706

2 Wisconsin Department of Natural Resources, 1350 Fermite Drive, Monona, Wl

53716

3 Corresponding author: [email protected]

Keywords: phosphorus, internal loading, physical conditions, eutrophication, climate

change

Status: In preparation for submission

Page 153: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

142 Abstract

Effects of climate warming are apparent in lake ecosystems evidenced by

rising water temperatures and shorter duration in ice cover. Modeling exercises,

studies of anomalous weather years, and examination of lakes across latitudinal

gradients suggest that changes in temperature will result in changes in the physical

structure of lakes, nutrient cycling, phenology, water quality and habitat availability.

Yet, few studies can empirically document the effects of warming trends on physical,

chemical and biological aspects of the system. Here we present a long-term record

from Lake Mendota of changes in ice cover, physical characteristics, phosphorus

dynamics and water quality metrics that provides evidence of the link between

changes in chemical and biological variables and changes in the physical

characteristics of the lake. We also examine the effects of changes in physical and

chemical variables on the general water quality characteristics of the lake. Temporal

trends show a decline in ice cover and correlated increases in strength and duration

of stratification. At the same time, we observed significant increases in phosphorus

concentrations and anoxic conditions in the hypolimnion. The trends in phosphorus

and oxygen conditions were likely due to a common relationship with increasing

stability and length of stratification over time. Change in the physical and chemical

features of the lake were tied to changes in water quality in that we also observed a

significant improvement in water clarity over time, and a significant relationship

between mean phosphorus concentrations in the epilimnion and stability of the water

column. Thus, our study provides empirical evidence of changing physical conditions

Page 154: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

143 in a deep dimictic lake and demonstrates that these physical changes have

implications for the distribution of solutes and the trajectory of water quality in lakes.

Page 155: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

144 Introduction

The effects of global warming are evident in lake ecosystems and lakes may

serve as a harbinger of further change at local and regional scales. Records of the

duration of ice cover show consistent and significant trends toward earlier ice break

up and later freeze dates across North America (Magnuson et al. 2000, Jensen et al.

2007). Also, recent lake ice cover records (since 1975) show an increasing rate of

change compared to historical changes in ice duration (Jensen et al. 2007). Changes

in the duration of ice cover are anticipated to have dramatic effects on stratification

and mixing patterns, on phenology and seasonal succession, on evaporation and

hydrology, and on the behavior and distribution of organisms that depend on these

systems.

Potential effects of climate change on aquatic ecosystems have been largely

explored through modeling exercises (Magnuson et al. 1997, Elliot et al. 2005).

Simulations of water temperature and dissolved oxygen characteristics in lakes of

different morphometries suggest that surface water temperature and the duration of

stratification can be expected to increase along with a lengthening of the period of

hypolimnetic anoxia (Stefan et al. 1996, DeStasio et al. 1996). From projected

changes in temperature and oxygen characteristics we can infer changes in the

distribution of fish species and biotic interactions within lakes (DeStasio et al. 1996,

Magnuson et al. 1997). Other modeling studies suggest likely changes in the timing

and species succession associated with the spring phytoplankton bloom (Elliot et al.

2005)

Page 156: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

145 In addition to modeling exercises, some studies have been able to use

anomalous weather years (Jankowski et al. 2006, Wilhelm and Adrian 2008) or a

latitudinal gradient (Weyhenmeyer 2004, Kosten et al. 2009) in lakes to explore the

potential influence of climate change on aquatic systems. These studies provide

empirical support for dramatic changes in thermal structure and concomitant changes

in oxygen distribution within lakes (Jankowski et al. 2006). Notably, Kosten et al.

(2009) found that an interaction between climate and nutrient concentrations

determined the vegetation structure within lakes across a climate gradient.

Although we have a number of models and long-term records of water

temperature (Livingstone 2003, Arhonditsis et al. 2004) that support the idea that

climate change will result in significant changes in the physical structure of lakes,

there are few studies that provide empirical evidence for the effects these physical

changes will have on the biology and chemistry of aquatic systems. Here we present

a thirty-year record from Lake Mendota of changes in ice cover, physical

characteristics, phosphorus dynamics and water quality metrics which provides

evidence for changes in internal processes resulting from changes in the physical

characteristics of the lake. We also examine the implications of changes in physical

and chemical features for the general water quality characteristics of the lake.

Materials and methods

All data were collected on Lake Mendota, a eutrophic dimictic lake in Madison,

Wisconsin (43°06' N, 89°25' W, 39.1 km2 surface area, 12.3 m mean depth). This

Page 157: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

146 lake has been monitored using consistent methods since 1975. The program

began under auspices of the Wisconsin Department of Natural Resources (WIDNR)

(1975 - 1995) and more recently has been monitored jointly by the WIDNR and the

North Temperate Lakes Long-Term Ecological Research program (NTL-LTER,

http://lter.limnology.wisc.edu). We compiled information on ice duration, water

temperatures, Secchi depth, phosphorus loading and dissolved oxygen and

phosphorus concentrations within the lake. We used these data to calculate metrics

related to lake stability, phosphorus dynamics, annual and summer-time phosphorus

mass balances, and the duration of stratification and anoxia.

Metrics related to the physical conditions of the lake included a metric related

to the duration of anoxia as a measure of the number of days that the lake was

strongly stratified. In each year we documented the first day on which oxygen

concentration at any depth was < 1.0 mg L"1 and documented the final day of the

summer on which anoxic conditions were observed. The difference between these

dates served as a proxy for the length of stratification. We also calculated the mean

Schmidt stability index for the lake over the summer. This calculation is a measure of

the total amount of energy that it would theoretically take to mix the entire lake

(Imberger and Paterson 1990). This metric is calculated as:

SS = -J-2(z-z*)(Pz-p*)AzAz A o

Where: SS = Schmidt stability (g cm"1)

A0 = surface area of the lake z = depth of interest

Page 158: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

147 z* = depth of mean density pz = water density at depth z p* = mean water density Az = area of lake at depth z Az = depth interval of calculation (1m)

Total phosphorus (P) samples were collected in acid-washed triple-rinsed

containers at 2 - 4 m intervals along the depth profile of the lake, and were generally

collected on a monthly basis. Phosphorus profiles were consistently collected in mid-

April and late October or early November in order to capture the P concentration in

the lake during spring and fall mixis. Temperature and dissolved oxygen

concentrations were recorded at every meter along a depth profile, and were

sampled every two weeks. All samples and measurements were collected at the

deepest point (~ 24 m) near the center of the lake. TP samples were preserved using

Optima HCI (1 N), and later analyzed using spectrophotometric techniques following

persulfate digestion, according to the most recent standard methods at the time

(APHAetal., 1995).

Details of methods related to P input to and export from the lake have been

reported previously in other studies (Lathrop et al. 1998, Lathrop 2007 and Carpenter

and Lathrop 2008). Briefly, there are four streams and two storm water inlets entering

Lake Mendota. Two streams (Pheasant Branch and Yahara River) and one storm

water inflow (Spring Harbor) are continuously monitored for hydrologic and chemical

inputs into the lake by the United States Geological Survey (USGS). Loading was

determined for these three inlets and these data were used to infer loading from other

inlets based on previous estimates of the relative load entering the lake from each

Page 159: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

source (Lathrop et al., 1998). Outflow of P was determined using the hydrologic

outflow of the lake multiplied by the mean epilimnetic P concentration.

Annual estimates of the phosphorus budget were based on calculations of the

cumulative P load and P outflow from the lake between April 16th of one year and

April 15th of the next. Annual estimates of load combined with hydraulic conditions for

the water-year were used as input variables for prediction of April P concentrations in

the spring based on the Vollenweider model (Vollenweider 1976):

[P] = (Lp/qs)*( L - ) l + Vz/qs

Where: [P] = mean P concentration (mg L"1) Lp = annual P load (g) qs = annual hydraulic load (m3) z = mean depth (m)

Additionally, we calculated a budget for the changes in P during the summer of

each year. Calculation of the P budget for the summer was based on observed April

(PA) and November (PN) P mass. Inputs and outputs from the system were quantified

using monthly estimates of P load and P outflow over the months between April and

November (this period will be referred to as the "summer"). Between April and

November of a given year, mass balance can be used to estimate November P mass

as follows:

PN = PA + SUMMLOAD - SUMMOUT + Netjlux

where PA and PN are P mass in April and November respectively, SUMMLOAD and

SUMMOUT are load and outflow, respectively, between April and November, and

Page 160: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

149 Net_flux is the difference between sediment release and sedimentation between

April and November, such that NeMlux is positive if recycling exceeds

sedimentation, and negative if sedimentation exceeds recycling. All budget terms

except NeMlux were measured directly. We estimated NeMlux by rearranging the

equation as:

Net.flux = (PN + SUMM0UT) - (PA + SUMML0AD)

In many lakes with anoxic hypolimnia, the rate of release of P from the

sediments is associated with the extent and duration of anoxic conditions due to the

release of P from chemical complexes in lake sediments under reducing conditions

(Nurnberg 1984, Nurnberg 1987). Therefore, the Anoxic Factor (AF), an index of

anoxic conditions, was explored as a potential predictor of hypolimnetic P conditions

in our long-term study. The AF is a measure of the number of days during the

summer that an area equivalent to the surface area of the lake is overlain by anoxic

water (Nurnberg 1987). The AF was calculated based on periodic oxygen profiles

collected between April and November in each year using the following:

A C _ Y (duration*anoxic_area) ^ lake_area

Where: AF = Anoxic Factor (days) Duration = number of days between oxygen profiles Anoxic_area = sediment surface area (m2) exposed to anoxic conditions (DO < 1.0 mg L"1) between sampling events

Page 161: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

150 Lake_area = total surface area of the lake (m2)

Other variables of interest included the rate of decline in dissolved oxygen

(DO), the rate of accumulation of P, and the P concentration observed in the

hypolimion in each September. For determination of the decline of DO and

accumulation of P, we quantified the total mass of DO and P in the hypolimnion,

which for this purpose was defined as the depths between 14 m and 24 m in depth.

As the mean thermocline depth is typically between 8 - 12 m, this definition of the

hypolimnion avoids incidental inclusion of water from the mixed layer. We

documented the DO or P in the hypolimnion at a number of sampling points during

the summer and used the slope of the mass over time as a measure of the rate of

DO decline or P accumulation. We excluded summers in which there were fewer than

3 observations of the P or DO profiles during the summer. Also, in order to identify a

reasonable slope, we limited our observation of DO profiles to the period between

May 1st and July 15th, the period during which oxygen concentrations in the

hypolimnion were observed to be changing. Similarly, observation of P mass in the

hypolimnion was limited to the period between April 15th and September 1st, the

period during which a clear trend in hypolimnetic P mass could be identified. The

maximum P concentration was observed at an index depth and date to facilitate

comparison across years.

I used the P concentration at 20 m depth for the sampling date in closest

proximity to September 1st (all sample dates fell between 25 August and 7

Page 162: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

151 September). An examination of the September P concentration versus the day of

year the sample was collected showed that the small variation in the timing of sample

collection did not affect the concentration of P observed.

All variables were normalized to a common annual time step based on the

mean of observed values. Analyses of the relationships among variables were

conducted using Pearson's Correlation, with an alpha significance of 0.05. All

analyses were completed using the R statistical software program (http://www.r-

project.org/).

Results

Metrics related to ice cover and the physical structure of the lake showed

notable temporal trends over the approximately 30-year period of record. Ice duration

showed a significant decline over time (p = 0.046, r = -0.37) (Figure 1a). Based on a

five-year average, the duration of ice cover has declined from a mean of 109 days

per winter to approximately 96 days of ice cover. The ice cover duration in 2001 (21

days) was identified as an outlier using Grubb's test, so this value was not included in

the reported five-year average.

The duration of stratification increased significantly over the last thirty years (p

= 0.046, r = 0.37) resulting in an increase in mean length of stratification from

approximately 122 to 139 days (again, based on a five-year average) (Figure 1b).

The Schmidt stability index also provided a signal that the physical structure of the

Page 163: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

152 lake has changed over time, yet this temporal trend was marginally significant (p

= 0.053, r = 0.35) (Figure 1c).

Some phosphorus (P) metrics have changed significantly over time. April P

mass and the mean concentration of P in the epilimnion, while temporally variable,

show no consistent trend over the period of record (Figure 2a and 2b, respectively).

On the other hand, the rate of P accumulation in the hypolimnion and the September

concentration of P in the hypolimnion have both increased notably since the late

1970's (p = 0.003, r = 0.57; p < 0.001, r = 0.82, respectively) (Figure 2c and 2d,

respectively).

We found evidence that the observed changes in hypolimnetic P dynamics are

related to changes in physical characteristics of the lake. The concentration of P in

the hypolimnion in September was positively correlated with the duration of

stratification (p = 0.005, r = 0.54) (Figure 3a). Yet, there was no significant

relationship between the September P concentration and the date of onset of

stratification (p = 0.22, r = -0.26). The rate at which P accumulated in the hypolimnion

was also weakly related to the duration of ice cover, but this trend was not statistically

significant at an a = 0.05 (p = 0.061, r = -0.38) (Figure 3b).

Hypolimnetic oxygen metrics also changed over thirty years and these

changes were associated with the strength and duration of stratification. The Anoxic

Factor increased significantly over time (p = 0.029, r = 0.41) (Figure 4) and these

changes were correlated with changes in the Schmidt stability index (p = 0.007, r =

0.49) (Figure 5a). Similarly, the rate of dissolved oxygen decline in the hypolimnion

Page 164: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

153 was negatively correlated with the date of onset of stratification (p = 0.009, r =

0.48). Therefore, when the lake stratifies early in the season, dissolved oxygen is lost

from the hypolimnion at a faster rate (Figure 5b). The oxygen metrics, however, were

not correlated with changes in P in the hypolimnion (rate of P accumulation or

September P concentration) (p > 0.05, r <0.2, in all cases). This indicates that the

temporal synchrony between P and oxygen conditions in the hypolimnion was driven

by a common relationship with increasing stability and length of stratification over

time.

These changes in the physical structure of the lake along with changes in the

phosphorus and oxygen dynamics in the hypolimnion have resulted in significant

changes in the water quality characteristics of the lake. The mean summer Secchi

depth reading has improved significantly over the period of study (p = 0.004, r = 0.52)

(Figure 6). The improvement in water clarity is positively related to an increase in the

stability of the water column (p = 0.016, r = 0.44) (Figure 7a). Additionally, the

stability of the water column is correlated with the mean P concentration in the

epilimnion (p = 0.015, r = -0.46) (Figure 7b). The mean epilimnetic P concentration

tends to be lower during summers with a high Schmidt stability index.

Given that we observed a significant temporal trend in characteristics

associated with internal loading (rate of P accumulation in the hypolimnion and

increasing late-season P concentrations in the hypolimnion) we wanted to explore the

potential effects of these changes on the annual phosphorus dynamics of the lake.

The summer season was bracketed by measurements of the P mass in the lake in

Page 165: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

154 April and November, during spring and fall turnover (Figure 8a). In many cases,

fall and spring values were quite similar and this relationship was reflected by a

significant positive correlation (r = 0.64, p <0.001). Generally, the PN was greater than

PA, but there were notable exceptions in 1993 and 1998. Summertime P loads

(SUMMLOAD) were quite high in some years, and peaks in summertime P loading

appeared more frequent in recent years (Figure 8). Loss of P from the outlet

(SUMMOUT) of the lake was relatively constant, with the exception of 1993 when an

unusually high summertime loading rate was associated with an unusually high

export of phosphorus from the lake. SUMMLOAD consistently exceeded SUMMOUT

(Figure 8b).

The balance of these input and output components was summarized as an

estimate of Net_Flux within the lake (Figure 8c). Sedimentation is known to be a large

component of the P flux within this eutrophic lake (Sonzogni et al. 1976, Soranno et

al. 1997), so it is not surprising that in most years the downward flux of P via

sedimentation outweighs the upward flux of internal loading from the sediments, thus

the value of Net_Flux is negative. Conversely, years in which internal loading

outweighs sedimentation are relatively rare. But, because positive peaks in the

Net_Flux highlight instances in which the net balance is in strongly in favor of one

component of the vertical flux, these peaks in net internal loading are particularly

notable.

While there is no apparent temporal trend in the Net_Flux (Figure 8c), which

serves as an index of net internal load, The residual of Net_Flux (after regression

Page 166: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

155 against the mean P concentration of the previous year) was positively correlated

with Paccum (p = 0.03, r = 0.45) (Figure 9). Thus, phosphorus in the hypolimnion

generally accumulated faster in years when sedimentation was relatively low. There

appeared to be a saturation of the relationship between net flux and P accumulation,

such that at high accumulation rates we didn't always see a further increase in the

net internal load. However, this relationship was difficult to interpret because the net

flux represents the relative balance between sedimentation and internal load. It is

possible that in summers with high P accumulation, there may also be higher rates of

sedimentation.

We explored whether any of the metrics related to internal loading could

explain residual variation associated with the Vollenweider model. The P load to Lake

Mendota has been variable, but shows no significant temporal trend (Figure 9a).

Similarly, the April P mass and associated residual error from the Vollenweider model

(Verr) show no significant temporal trend (Figure 9 b and 9c). Variability in Verr does

seem to have decreased in recent years (Figure 9c).

The rate of accumulation of P in the hypolimnion, the maximum concentration

of P observed, and the net flux metric from the previous year were not significantly

related to the residual errors of the Vollenweider model (Verr). Instead, a display of the

autocorrelation function for April P mass suggests that there is a 1-year time lag in

the mass of P in the lake in April, indicating that some variation in observed April P

mass can be explained by the previous April P mass in the previous year (Figure 11).

In fact, we found a simple linear regression model based on April P mass from the

Page 167: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

156 previous year and the total P load from the previous year could explain the April P

mass in the year of interest quite well (p <0.001, R2 = 0.67, AIC = 527.6). In this case,

adding net flux from the previous year to the model resulted in a marginally significant

improvement in model fit (coefficient p-value = 0.064, AIC = 525.4). Thus, there is

some support for the idea that release of phosphorus from the sediments (as a "new"

source of P) contributes to the April P mass in the lake.

We did observe a relationship between the rate of P accumulation in the

hypolimnion and the residual variation around the Vollenweider relationship. The

residuals from the Vollenweider model (an index of Spring P conditions) were

correlated with the net flux metric for the following summer (p = 0.05, r = 0.39) (Figure

12). Thus, when April P concentrations were higher than predicted by the

Vollenweider model (i.e. a negative residual) it was most likely that there would be

net sedimentation during the summer. Net recycling was most prevalent in years

when the Vollenweider model predictions were greater than the observed April P

concentrations.

Discussion

Changes in lake stability and length of stratification are associated with

changes in the duration of ice cover. Long-term records document a decrease in ice

cover on Lake Mendota and many lakes worldwide (Magnuson et al. 2000). Scenario

modeling exercises corroborate the relationship between the stability and duration of

stratification in lakes and climate change (Hondzo and Stefan 1992). Other long-term

Page 168: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

157 studies indicate that surface waters warm at a higher rate compared to bottom

waters, thus inducing changes in physical stability of the water column and duration

of stratification (Livingstone 2003). Our study provides empirical evidence of

changing physical conditions in a deep dimictic lake and demonstrates that these

physical changes have implications for the distribution of solutes and the trajectory of

water quality in lakes.

Changing physical conditions in the lake played a dominant role in determining

the distribution of P and DO during the summer with implications for water quality. As

length of stratification increases, we might expect P to build up in the hypolimnion

over a longer period of time (Nurnberg 1998). However, our data suggest this is not

the sole reason for the increasing trends in maximum P concentrations in the

hypolimnion because Psept was consistently measured around September 1st in all

years and the magnitude of Psept was not significantly correlated with the date of

onset of stratification. Instead we think that the increase in Psept may be partly

attributed to a lower vertical flux rate of P to the epilimnion and higher rates of

sedimentation from the epilimnion. During years with relatively low stability, the

vertical eddy diffusivity would be higher, thus allowing higher rates of vertical diffusion

through the hypolimnion and potentially reducing sedimentation (Stuaffer 1992). Thus

changes in the P concentration in the hypolimnion of Lake Mendota are likely due to

an increase in the period over which P may accumulate, increased sedimentation,

and to effective containment of P within the hypolimnion.

Page 169: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

158 The importance of physical conditions in vertical solute transport and water

quality are in line with previous studies that show entrainment events and algal

blooms are tied to changing weather dynamics at a daily scale (Soranno 1997,

Stauffer 1992). Similarly, Lathrop and others (1999) found that interannual variation in

Secchi depth could be explained by lake stability, along with Daphnia grazing and P

availability. The effects of climate change, however, are not likely to have positive

effects on water quality across lakes. An increase in strength and duration of

stratification could have profound repercussions in shallow lakes with intermittent

mixing. Polymictic lakes are often prone to short-term stratification, and as such may

experience bouts of anoxia and internal loading (Riley and Prepas 1984, Kallio 1994,

Wilhelm and Adrian 2008). If climate change can be expected to increase the

incidence of these internal loading episodes, water clarity would likely decline.

Net_Flux was positively correlated with the rate of P accumulation in the

hypolimnion (PaCcum). It follows that PaCcum rates are relatively low when net

sedimentation is high, which may imply that P accumulating in the hypolimnion is not

derived from sedimentation in the same year. Instead, correlation between the

September P concentration at 20 m (Psept) and the mean epilimnetic P concentration

(Pepi) of the previous year lend support to the idea that P accumulating in the

hypolimnion is due to recycling of P from the previous year.

A short lag time between deposition of P in the hypolimnion and

remineralization is further supported by the fact that none of the three metrics of

internal loading (rate of P accumulation, September P concentration at 20 m, and the

Page 170: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

159 net flux metric from the previous summer) could explain residual variation around

the Vollenweider model predictions of April P mass. Additionally, the 1-year time lag

in the autocorrelation function for April P mass suggests the memory term (via the

mechanism of internal loading) for the system is relatively short-lived.

This interpretation is illustrated well by observations during the period between

1987 and 1992, which followed a drought in 1987-1988. During this period external

loads were relatively low and consistent, if internal loading were a dominant process,

we would expect to see evidence of net P release particularly during the summer

months. However, we see that the November phosphorus mass very closely tracks

the phosphorus mass observed in April, and the difference between the two can very

nearly be explained by the difference between summer-time external load and P

outflow (Figure 8). Thus, a number of lines of evidence from this long-term study

suggest that mean P conditions in Lake Mendota are predominantly controlled by

external loading to the system. This finding supports previous studies following the

diversion of wasterwaters from Lake Mendota which suggested that, overall, the

sediments of the lake act as a sink rather than source of phosphorus to the system

(Sonzogni and Lee 1974).

Changes in climatic patterns are already having noticeable effects on lake

ecosystems. The resultant changes in water temperature and physical stability of a

lake have the power to exact changes in oxygen dynamics and internal nutrient

cycling. Using a long-term data set we provide empirical evidence for changes in the

physical characteristics for Lake Mendota and show that changes in oxygen,

Page 171: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

160 phosphorus and water quality have occurred at the same time. While there have

been significant changes in internal phosphorus dynamics in the lake, the error

around the Volllenweider model (which predicts April P conditions) could not be

explained by these changes in internal P loading. Thus, P dynamics for the system

remain driven by external loading, and reductions of external loading, not further

climate change, are the key to restoration of water quality.

Page 172: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

References 161

Arhonditsis, G. B., M. T. Brett, C. L. DeGasperi, and D. E. Schindler. 2004. Effects of climate variability on the thermal properties of Lake Washington. Limnology and Oceanography 49:256-270.

Bostrom, B., G. Persson, and B. Broberg. 1988. Bioavailability of Different Phosphorus Forms in Fresh-Water Systems. Hydrobiologia 170:133-155.

Caraco, N. F., J. J. Cole, and G. E. Likens. 1989. Evidence for Sulfate-Controlled Phosphorus Release from Sediments of Aquatic Systems. Nature 341: 316-318.

Carpenter, S.R., and R. C. Lathrop. 2008. Probabilistic estimate of a threshold for eutrophication. Ecosystems 11: 601-613.

DeStasio Jr., B. T., D. K. Hill, J. M. Kleinhaus, N. P. Nibbelink, and J. J. Magnuson. 1996. Potential effects of global climate change on small north-temperate lakes: Physics, fish and phytoplankton. Limnology and Oceanography 41:1136-1149.

Elliot, J. A., S. J., Thackeray, C. Huntingford, and R. G. Jones. 2005. Combining a regional climate model with a phytoplankton community model to predict future changes in phytoplankton in lakes. Freshwater Biology 50:1404-1411.

Gachter, R., and B. Muller. 2003. Why the phosphorus retention of lakes does not necessarily depend on the oxygen supply to their sediment surface. Limnology and Oceanography 48: 929-933.

Golterman, H. L 1977. Interactions between sediments and fresh water: proceedings of an international symposium held at Amsterdam, the Netherlands, September 6-10, 1976. W. Junk.

—. 2001. Phosphate release from anoxic sediments or 'What did Mortimer really write?' Hydrobiologia 450: 99-106.

Graneli, W. 1999. Internal phosphorus loading in Lake Ringsjon. Hydrobiologia 404: 19-26.

Gibson, C. E., G. X. Wang, R. H. Foy, and S. D. Lennox. 2001. The importance of catchment and lake processes in the phosphorus budget of a large lake. Chemosphere 42: 215-220.

Holdren, G. C, and D. E. Armstrong. 1980. Factors Affecting Phosphorus Release from Intact Lake Sediment Cores. Environmental Science & Technology 14: 79-87.

Page 173: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

162

Hondzo, M., and H. G. Stefan. 1993. Regional Water Temperature Characteristics of Lakes Subjected to Climate-Change. Climatic Change 24:187-211.

Imberger, J., and J. C. Patterson. 1990. Physical Limnology. Advances in Applied Mechanics 27: 303-475.

Jeppesen, E., M. Sondergaard, J. P. Jensen, E. Mortensen, A. M. Hansen, and T. Jorgensen. 1998. Cascading trophic interactions from fish to bacteria and nutrients after reduced sewage loading: An 18-year study of a shallow hypertrophic lake. Ecosystems 1: 250-267.

Jeppesen, E. and others 2005. Lake responses to reduced nutrient loading - an analysis of contemporary long-term data from 35 case studies. Freshwater Biol 50: 1747-1771.

Kallio, K. 1994. Effect of Summer Weather on Internal Loading and Chlorophyll-a in a Shallow Lake - a Modeling Approach. Hydrobiologia 276:371-378.

Lathrop, R. C, S. R. Carpenter, and D. M. Robertson. 1999. Summer water clarity responses to phosphorus, Daphnia grazing, and internal mixing in Lake Mendota. Limnology and Oceanography 44:137-146.

Lathrop, R. C, S. R. Carpenter, C. A. Stow, P. A. Soranno, and J. C. Panuska. 1998. Phosphorus loading reductions needed to control blue-green algal blooms in Lake Mendota. Canadian Journal of Fisheries and Aquatic Sciences 55:1169 -1178.

Livingstone, D. M. 2003. Impact of secular climate change on the thermal structure of a large temperate Central European lake. Climate Change 57:205-225.

Magnuson, J. J., D. M. Robertson, B. J. Benson, R. H. Wynne, D. M. Livingstone, T. Arai, R. A. Assel, R. G. Barry, V. Card, E. Kuusisto, N. G. Granin, T. D. Prowse, K. M. Stewart, and V. Vuglinski. 2000. Historical trends in lake and river ice cover in the Northern Hemisphere. Science 289:1743-1746.

Magnuson, J. J., K. E. Webster, R. A. Assel, C. J. Bowser, P. J. Dillon, J. G. Eaton, H. E. Evans, E. J. Fee, R. I. Hall, L. R. Mortsch, D. W. Schindler, and F. H. Quinn. 1997. Potential effects of climate change on aquatic systems: Laurentian Great Lakes and Precambrian Shield Region. Hydrological Processes 11:825-871.

Marsden, M. W. 1989. Lake Restoration by Reducing External Phosphorus Loading -the Influence of Sediment Phosphorus Release. Freshwater Biol 21:139-162.

Page 174: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

163 Moosmann, L, R. Gachter, B. Muller, and A. Wuest. 2006. Is phosphorus retention in autochthonous lake sediments controlled by oxygen or phosphorus? Limnology and Oceanography 51: 763-771.

Mortimer, C. H. 1941. The exchange of dissolved substances between mud and water in lakes. Journal of Ecology 29: 280-329.

Nurnberg, G. K. 1984. The Prediction of Internal Phosphorus Load in Lakes with Anoxic Hypolimnia. Limnology and Oceanography 29:111-124.

—. 1987. A Comparison of Internal Phosphorus Loads in Lakes with Anoxic Hypolimnia - Laboratory Incubation Versus Insitu Hypolimnetic Phosphorus Accumulation. Limnology and Oceanography 32:1160-1164.

—. 1998. Prediction of annual and seasonal phosphorus concentrations in stratified and polymictic lakes. Limnology and Oceanography 43:1544-1552.

Riley, E. T., and E. E. Prepas. 1984. Role of Internal Phosphorus Loading in 2 Shallow, Productive Lakes in Alberta, Canada. Canadian Journal of Fisheries and Aquatic Sciences 41: 845-855.

Sondergaard, M., P. Kristensen, and E. Jeppesen. 1993. 8 Years of Internal Phosphorus Loading and Changes in the Sediment Phosphorus Profile of Lake Sobygaard, Denmark. Hydrobiologia 253: 345-356.

Sondergaard, M., J. P. Jensen, and E. Jeppesen. 2001. Retention and internal loading of phosphorus in shallow, eutrophic lakes. The Scientific World 1: 427-442.

Sonzogni, W. C, and G. F. Lee. 1974. Diversion of Wastewaters from Madison Lakes. Journal of the Environmental Engineering Division-Asce 100:153-170.

Sonzogni, W. C, P. C. Uttormark, and G. F. Lee. 1976. Phosphorus Residence Time Model - Theory and Application. Water Research 10: 429-435.

Soranno, P. A. 1997. Factors affecting the timing of surface scums and epilimnetic blooms of blue-green algae in a eutrophic lake. Canadian Journal of Fisheries and Aquatic Sciences 54:1965-1975.

Soranno, P. A., S. R. Carpenter, and R. C. Lathrop. 1997. Internal phosphorus loading in Lake Mendota: response to external loads and weather. Canadian Journal of Fisheries and Aquatic Sciences 54:1883-1893.

Page 175: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

164 Stauffer, R. E. 1985. Relationships between Phosphorus Loading and Trophic State in Calcareous Lakes of Southeast Wisconsin. Limnology and Oceanography 30:123-145.

—. 1987. Vertical Nutrient Transport and Its Effects on Epilimnetic Phosphorus in 4 Calcareous Lakes. Hydrobiologia 154: 87-102.

—. 1992. Efficient Estimation of Temperature Distribution, Heat-Storage, Thermocline Migration and Vertical Eddy Conductivities in Stratified Lakes. Freshwater Biol 27: 307-326.

Stefan, H. G., M. Hondzo, X. Fang, J. G. Eaton, and J.H. McCormick. 1996. Simulated long-term temperature and dissolved oxygen characteristics of lakes in the north-central United States and associated fish habitat limits. Limnology and Oceanography 41:1124-1135.

Van der Molen, D. T., and P. C. M. Boers. 1994. Influence of Internal Loading on Phosphorus Concentration in Shallow Lakes before and after Reduction of the External Loading. Hydrobiologia 276: 379-389.

Vollenweider, R. A. 1976. Advances in defining critical loading levels for phosphorus in lake eutrophication. Mem. 1st. Ital. Idrobiol 33: 53-83.

Wilhelm, S., and R. Adrian. 2008. Impact of summer warming on the thermal characteristics of a polymictic lake and consequences for oxygen, nutrients and phytoplankton. Freshwater Biology 53:226-237.

Page 176: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

165

80

100

120

1 1

1

40

60

o _

• • • •

(a) p = 0.046 r = -0 .37

i 1 i i

• • •

• •

• •

• •

1 1 •

• •

• •

1 1975 1980 1985 1990

Year

1995 2000 2005

!t= o

o r-- -

150

o CO -

(b) p = 0.046 r = 0 . 3 7

• • • •

1975 1980

• • • • •

1985 1990

Year

1995 — I —

2000 2005

X CD "O C *"-?r

01 V)

• D

E f-

a m (0

o u> m

o in

o

(c) p = 0.053 r = 0.35

• • •

• •

• • •

• •

1975 1980 1985 — I 1 —

1990 1995

Year

— i —

2000 2005

Figure 1. Temporal trends in ice cover (a), the duration of stratification (b), and the Schmidt stability index (c). Ice cover and duration of stratification show significant trends over time (r = Pearson's Correlation coefficient, p = p-value). Schmidt stability index shows a temporal trend which is marginally significant at an a = 0.05. The ice cover value for 2001 (21 days) was identified as an outlier (Grubb's test)

Page 177: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

166

o o 8 CO

d E

a <

8"

o

u E

I a.

(a) p = 0.89 r = -0.03

• •

• • •

• • • • • •

• • I 1 t

• •

1

-

• • •

• 1 1

1975 1980 1985 1990 1995 2000 2005

(c) p = 0.003 r = 0.57

• •

I - ' 1

• • • •

• •

• • •

1 1 r | - 1

(b)

p=0.73 rm= -0.07

1 1 1 1 1 r 1975 1980 1985 1990 1995 2000 2005

E. E •" o d tM

«

XI O

E CD

(d) p< 0.001 r = 0.82

1975 1980 1985 1990 1995 2000 2005 ' i 1 1 1 1 1 r 1975 1980 1985 1990 1995 2000 2005

Figure 2. Temporal trends in phosphorus dynamics in Lake Mendota. There is no significant temporal trend in the mass of phosphorus in the lake in April of each year (a) or in the mean epilimnetic phosphorus concentration (b). The rate of phosphorus accumulation in the hypolimnion (c) and the phosphorus concentration at 20 m depth in September (d) both showed significant increases over time (r = Pearson's correlation coefficient, p = p-value).

Page 178: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

167

^ -Li O) E

• * — * *

E o CM Cfl

l i |

Q. • — ^

d> J3 E <D Q. d>

in (D o

lO in d

in »* b

in CO

(a) p = 0.005 r = 0.54

o

o

o *

1

o

o 1

o 8

o

CD

O

O

1

o

1

o

o

o

o

o

o

1 1

o

1

110 120 130 140 150 160 170 180

Duration of stratification (days)

T

T3

C .o " • * • -

m 3

E 8 A

Q. 1 -

o Q) 15 QC

1400

o o o ^"

o o CD

o o CM

-

o

I

20

I

40

o

o

I

60

o o

o o

o o

o o I

80

o

o

K i o

I

100

(b) p = 0.061 r = -0.38

o

a °

i 120

Ice duration (days)

Figure 3. Hypolimnetic P variables were related to changes in physical characteristics of the lake. The phosphorus concentration at 20 m depth in September was significantly correlated with the duration of stratification (a). There was a negative trend in the relationship between ice duration and the rate of P accumulation in the hypolimnion (b), but this relationship was not statistically significant (r = Pearson's Correlation coefficient, p = p-value).

Page 179: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

168

o

o

O lO

o

p = 0.029 r = 0.41

• •

1 1 1 1— 1975 1980 1985 1990 1995

Year

2000 2005

Figure 4. There was a significant increase in the Anoxic Factor over time (r = Pearson's Correlation coefficient, p = p-value).

Page 180: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

169

o U_ O "x o c <

o

o

O lO

O

(a) p = 0.007 r = 0.49

o o o'

o o

° iO O O

T T T T T T 350 400 450 500 550 600 650

Schmidt stability index

>» rt

• o i !_J U)

E <D _c "o <D

• o

O Q H —

o (D

-i—< CS

cc

lO CM d

o d

lO

d

o ^ o

o o o

o

o

° o° oo o 0

(b) p = 0.009 r = 0.48

8

150 160 170 180 190

Onset of stratification (day) Figure 5. Changes in the oxygen characteristics of the hypolimnion were significantly related to changes in the physical characteristics of the lake. The Anoxic Factor was positively correlated with the Schmidt stability index (a), while the rate of dissolved oxygen decline in the hypolimnion was significantly correlated with the date of onset of stratification (r = Pearson's Correlation coefficient, p = p-value).

Page 181: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

o o CD

CO i— CD E E CO

c ctf CD

in CO

Q.

•D 9 co

c\i

o c\i

p = 0.004 r = 0.52

• •

T 1975 1980 1985

— I 1 — 1990 1995

Year

2000 2005

Figure 6. The mean summer time Secchi depth reading has increased significantly over time (r = Pearson's Correlation coefficient, p = p-value).

Page 182: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

171

CO

O

CO

uo cvi

o CVJ

(a) p = 0.016 r = 0.44

• •

350 400 450 500 550 600 650

Schmidt stability index

—̂̂ 1

!_i O) E

• * — '

epi

CL

c CO CD

^^.

• *

T -

o

o 'r"i o

CD o

(b) p = 0.015 r = -0.46

350 400 450 500 550 600 650

Schmidt stability index

Figure 7. Changes in the Secchi depth are positively correlated with changes in the Schmidt stability index (a), while there is a negative relationship between the mean phosphorus concentration in the epilmnion and the Schmidt stability index (r = Pearson's Correlation coefficient, p = p-value).

Page 183: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

172

J*.

CO

tn

E

.o CD 3

E o

o

o + CD

o + CD

CD

O + CD

1975 1980 1985 1990 1995 2000 2005

o

o

CM

SUMMLOAD _ (b) SUMMQUT A .

1975 1980 1985 1990 1995 2000 2005

1985

Year Figure 8. Summer phosphorus metrics. Phosphorus mass in the lake at spring (April, PA) and fall (November, PN) mixis (a). Total phosphorus inputs (Load) and outputs (Outflow) during the period between April and November (b). The summer Net flux is a measure of the balance between inputs, outputs and the starting and ending P mass: Net flux = (PN + SUMMQUT) - (PA + SUMMLOAD).

Page 184: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

173

"o o o

<D

c g

•4—<

a •4—<

C (1) E (D W

-i—>

(D

X

•g

CD

cr

o o o o

o -

o o o o

o o o o I

p = 0.03 r = 0.45

• •

• •

• • •

• •

1 1

• • •

1

1

i I

1 200 400 600 800 1000 1400

, - 1 ' P accumulation (kg day )

Figure 9. There is a significant positive relationship between the rate of P accumulation in the hypolimnion during the summer and the Net_flux over that same summer. Rate of P accumulation is slightly higher in years when sedimentation is low.

Page 185: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

g (« O

c c <

174

1975 1980 1985 1990 1995 2000 2005

o d

fc O <B O

> d

o i

1975

Figure 10. Annual phosphorus metrics. Annual P load (a) represents the total external load that entered the system between April 16th of the previous year and April 15th of the year of interest. April P mass (b) represents the total P mass in the lake on April 15th of the year of interest. Panel (c) represents the error in the prediction of the April P concentration based on the Vollenweider model. When Verr is positive, observed P concentration < predictions; when Verr is negative, observations > predicted P concentration.

Page 186: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

175

Q. <

to W E

Q.

"5 o <

0 0.

5 1.

0

L_

I .

o

m d -i 1 1

(/) (3 E

00 O

(O O

d

(0

fir

*•— o u-O <

CM

d

o d

3 -• 8 10 2

T" 4

T " 6 8

1 10

Lag Lag

Figure 11. Autocorrelation function (ACF) (a) for the time series of P mass in Lake Mendota in April of each year. Note the significant autocorrelation (a) at a 1-year time lag. We also present the ACF of the first difference between successive April P mass observations (b).

Page 187: Long-term trends in aquatic pollutants: Chloride and ... · Long-term trends in aquatic pollutants: Chloride and phosphorus dynamics in lakes embedded in urban and agricultural watersheds

O) c 73 >* o Q) *-

z ^_^

(kg

3

o o o m ^

o o

CD

<D

Z

c o •4—1

(0 -•—• c (1) E

edi

V)

o (/}

rt u ;u "</) d) cr

o o o in i

000

m

i 1 1 r -0.04 0.00 0.02 0.04 0.06

V, err

Figure 12. There was a marginally significant positive relationship between the residuals of the Vollenweider model (Verr) and the residuals of the Net Flux metric (after regression against mean P in the epilimnion).