185
AN ABSTRACT OF THE DISSERTATION OF Sarah L. Close for the degree of Doctor of Philosophy in Zoology presented on January 17, 2014. Title: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of Rocky Intertidal Macrophyte Assemblages Abstract approved: _____________________________________________________________________ Bruce A. Menge The influence of the physical environment on organisms has long been a subject of ecological research. But, the complex drivers of environmental variation, and the multiple scales at which this can occur, make studying this topic a difficult challenge. In rocky intertidal habitats, oceanographic- and climate-scale variability influence benthic communities primarily via changes in temperature, propagule delivery, and nutrient availability. Nutrient availability especially is thought to influence the diverse and productive macrophyte assemblages in these communities. While bottom-up regulation of communities provisioned by resource subsidies is known to play a role in community dynamics in this system, we are still developing an understanding of how nutrients influence macrophytes. In this dissertation, I investigated the environmental drivers of intertidal macrophyte ecology at both the biogeographic (regional) and the local scale. Using data spanning 10 years and over 900 km of coastline within the California Current Large

The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

AN ABSTRACT OF THE DISSERTATION OF

Sarah L. Close for the degree of Doctor of Philosophy in Zoology presented on January

17, 2014.

Title: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and

Functioning of Rocky Intertidal Macrophyte Assemblages

Abstract approved:

_____________________________________________________________________

Bruce A. Menge

The influence of the physical environment on organisms has long been a subject

of ecological research. But, the complex drivers of environmental variation, and the

multiple scales at which this can occur, make studying this topic a difficult challenge. In

rocky intertidal habitats, oceanographic- and climate-scale variability influence benthic

communities primarily via changes in temperature, propagule delivery, and nutrient

availability. Nutrient availability especially is thought to influence the diverse and

productive macrophyte assemblages in these communities. While bottom-up regulation of

communities provisioned by resource subsidies is known to play a role in community

dynamics in this system, we are still developing an understanding of how nutrients

influence macrophytes.

In this dissertation, I investigated the environmental drivers of intertidal

macrophyte ecology at both the biogeographic (regional) and the local scale. Using data

spanning 10 years and over 900 km of coastline within the California Current Large

Page 2: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

Marine Ecosystem (CCLME), I first evaluated the contributions of nutrient availability,

upwelling, spatiotemporal variation, and climate variability to macrophyte tissue nutrient

content. The insights gained from this investigation show a strong influence of nutrient

availability on nutrient status, although that relationship varied among species. I also

observed remarkably similar interannual variability in tissue nutrient content, which is

consistent with regional and basin-scale oceanographic processes. My results in Chapter

2 point to a role of nutrient availability in modulating macrophyte nutrient content which

is consistent with regional forcing in the CCLME.

In Chapter 3 I asked how nutrient and light availability drive growth, and how this

relates to nutrient content. Combining insights from Chapter 2 on macrophyte nutrient

content with data on growth rates in two dominant low intertidal species, Saccharina

sessilis and Phyllospadix scouleri, I found contrasting drivers of growth in the two

species. Growth in S. sessilis was positively related to nutrient availability, consistent

with our expectations, but negatively related to light availability, confounding

predictions. Growth in P. scouleri, on the other hand, showed more complicated

dynamics in relation to nutrients and light, suggesting that the abundance of these

resources modulates the growth response. These patterns nonetheless suggest that growth

in these species is strongly influenced by the upwelling process, although the

mechanisms behind such influences may vary.

Finally, in Chapter 4 I approached the question of how the lower limits of one

high zone intertidal species are set, that of Fucus distichus, to determine local scale

environmental drivers of macrophyte distribution. Through a field experiment, I

Page 3: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

manipulated F. distichus across an intertidal gradient and found that it had high mortality

at low intertidal elevations. Using a laboratory mesocosm experiment, I found that the

mechanism behind the elevational pattern appears to be the reduced light availability in

the low intertidal zone. Combined with the insights from Chapter 3, which suggests that

light influences macrophyte growth, the result that light could play a role in setting the

elevational limit in this species also underscores the importance of considering this

resource in studies of macrophyte ecology.

Page 4: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

© Copyright by Sarah L. Close

January 17, 2014

All Rights Reserved

Page 5: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and

Functioning of Rocky Intertidal Macrophyte Assemblages

by

Sarah L. Close

A DISSERTATION

submitted to

Oregon State University

in partial fulfillment of

the requirements for the

degree of

Doctor of Philosophy

Presented January 17, 2014

Commencement June 2014

Page 6: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

Doctor of Philosophy dissertation of Sarah L. Close presented on January 17, 2014.

APPROVED:

_____________________________________________________________________

Major Professor, representing Zoology

_____________________________________________________________________

Chair of the Department of Zoology

_____________________________________________________________________

Dean of the Graduate School

I understand that my dissertation will become part of the permanent collection of Oregon

State University libraries. My signature below authorizes the release of my dissertation to

any reader upon request.

_____________________________________________________________________

Sarah L. Close, Author

Page 7: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

ACKNOWLEDGEMENTS

The experiences I have had working towards my Ph.D. in Corvallis have been

incredible, and would not have happened without the help, encouragement, and support

of many, many people. I am indebted to my advisor, Bruce Menge for his incredible

support and advice over the years. His open door and ability to help me think critically

about my research have helped me to grow as a scientist. Days spent with Bruce in the

field, in Oregon and New Zealand, have been some of the most fun and educational

experiences of my time at OSU. Additionally, Jane Lubchenco co-advised me when I

first started my Ph.D. research and I am also grateful for her advice and continued

inspiration.

I am also extremely grateful to my graduate committee. Sally Hacker has been an

incredible mentor and collaborator. Her encouragement and advice have improved my

research and have broadened my perspectives. My conversations with Matt Bracken have

never failed to clarify my thinking on complicated subjects, and I greatly appreciate his

involvement on my committee. Sarah Henkel’s support and guidance have improved my

research and strengthened my ideas. Finally, I thank Brenda McComb for serving as my

Graduate Council Representative and truly being an advocate for me. To a person, my

committee has shaped my work and I am grateful for their contributions.

I cannot imagine these past 5 ½ years without the support of the amazing

LubMenge lab. From the beginning, my labmates have become like family: we push each

other, support each other, and celebrate with each other. Luis Vinueza, Joe Tyburczy, and

Dafne Eerkes-Medrano showed me the ropes and have always been nothing but

Page 8: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

encouraging. Conversations with Margot Hessing-Lewis are always refreshing and have

renewed my excitement in my research and made me think about it in new ways on

multiple occasions. I have been fortunate to share an office with Alison Iles for most of

my Ph.D. and I could not have asked for a better friend and officemate. I am grateful for

the support, friendship, and camaraderie of Jeremy Rose. Allie Barner, Liz Cerny-

Chipman, Chenchen Shen, Jessie Reimer, and Jenna Sullivan have all been incredible

friends and supporters throughout this process. Whether it is field trips to Portland,

escapes to the mountains, or quick chats in the hallway, my adventures with my labmates

have been some of the most valued experiences of my Ph.D. Other lab members as well

have been integral to my success. Kirsten Grorud-Colvert, Tarik Gouhier, Leigh Tait, and

Annaliese Hettinger have all been role models and mentors. I am indebted to lab manager

Jerod Sapp, technicians Gayle Murphy, Camryn Pennington, Megan Poole, Ruth Milston-

Clements, Becky Focht, Lindy Hunter, Tully Rohrer, Jonathan Robinson, Ryan Craig,

Angela Johnson, and Shawn Gerrity, and many interns, whose tireless work and good

humor have made field- and lab-work a joy. Much of this work was the result of research

that began long before I started my Ph.D., so I am grateful for the foresight and

stewardship of those who came before me. I also thank Cindy Kent and Kathleen Norris

for their support and patience throughout my Ph.D.

I am lucky to count myself among the members of the Zoology department

community. The faculty have been nothing but supportive. I thank the many friends

among Zoology graduate students who have helped me through this process, whether it is

answering mundane questions, providing support during difficult periods, or celebrating

Page 9: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

little victories. I have made many lifelong friends in this department. I especially thank

Orissa Moulton, Phoebe Zarnetske, Kate Boersma, Lindsay Biga, Paul Bradley, and Erin

Gorsich, among many, many others, for being amazing friends, confidants, and role

models. I am grateful for the support of our amazing administrative staff. Truly not

enough can be said about the work that Tara Bevandich, Torri Givigliano, Trudy Powell,

and Traci Durrell-Khalife put into making this department a wonderful place.

This work is also the result of productive and ongoing collaborations. Karina

Nielsen has challenged me and taught me new ways to think about my research and I am

grateful to her for her continued support and assistance. I also thank Francis Chan, who

has always helped me to see my research from different perspectives. Additionally, Tarik

Gouhier taught me to stand on my own two feet when it comes to statistics, and has

always been more than willing to provide a helping hand. Leigh Tait provided valuable

assistance in thinking about the role of light in the distribution of intertidal organisms,

and technical help in measuring photosynthesis in Chapter 4. I have had numerous

student workers, mostly volunteers, who have given their time to assist this research.

They are: Amandarose Kiger, Megan Atkinson, Meghan Beazley, Jonathan Robertson,

Kevin Krallman, Kristin Beem, Cassidy Huun, and Sean Harrison.

Finally, my family and friends have been a continued source of inspiration,

encouragement, and joy. My love for the ocean came at an early age and I am grateful for

the support of my parents, Wendy and David Close, who gave me every opportunity to

challenge myself and pursue my passions. They have always encouraged me beyond

measure. My sisters, Marian and Emma, have been my sidekicks my whole life. The

Page 10: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

support and friendship of my sisters and their families has fueled me through the ups and

downs of this process and made me laugh uncontrollably when I most needed it. I am

also incredibly grateful to the many friends who have shared adventures with me over the

years and whose support has helped me reach this point.

My dissertation research was funded by a variety of sources. I was supported on

teaching assistantships through the Biology program for multiple terms, for which I am

grateful. I also received generous research assistantship funding from Bruce for two

terms. I received a National Science Foundation Graduate Research Fellowship, the

Hannah T. Croasdale Fellowship from the Phycological Society of America, and three

awards from the Zoology Department Research Fund (ZoRF). I also received a Mamie

Markham Research Award from Hatfield Marine Science Center (HMSC), and I thank

HMSC for logistical and facilities support for the research conducted there (presented in

Chapter 4). I have received travel support from the Oregon State University Graduate

School and College of Science. The C:N and growth measurement studies were largely

funded by awards from the National Science Foundation, including NSF grant numbers

OCE 0726983, 1061530, 0727611, and 1061233 to BAM, FC, KJN, and SDH, NSF

grant number DEB 1050694 to BAM. Additional funding came from endowment funds

from the Wayne and Gladys Valley Foundation, grants from the A. W. Mellon

Foundation to BAM and Jane Lubchenco, and the David and Lucile Packard and Gordon

and Betty Moore Foundations for support of the Partnership for Interdisciplinary Studies

of Coastal Oceans (PISCO) consortium to BAM, JL, and lead PIs.

Page 11: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

CONTRIBUTION OF AUTHORS

Numerous collaborators and coauthors played an integral role in this work. Their

contributions are listed here. For Chapters 2 and 3, Francis Chan, Karina J. Nielsen, Sally

D. Hacker, and Bruce A. Menge all provided data, insight, and guidance and are included

as co-authors. For Chapter 4, Bruce A. Menge provided advice and assistance at all

stages of the research and is included as a co-author.

Page 12: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

TABLE OF CONTENTS

Page

Chapter 1 – General Introduction .................................................................................. 1

Chapter 2 – Organization out of chaos? Separating the effects of species, temporal

variation, climate, and spatial scales as determinants of intertidal macrophyte

nutrient dynamics ............................................................................................... 7

Abstract ................................................................................................................... 7

Introduction ............................................................................................................. 8

Methods ................................................................................................................. 14

Results ................................................................................................................... 20

Discussion ............................................................................................................. 26

Acknowledgements ............................................................................................... 35

Figures ................................................................................................................... 37

Tables .................................................................................................................... 45

Chapter 3 – Contrasting drivers of growth rate in two dominant rocky intertidal

macrophytes ..................................................................................................... 57

Abstract ................................................................................................................. 57

Introduction ........................................................................................................... 58

Methods ................................................................................................................. 64

Results ................................................................................................................... 69

Discussion ............................................................................................................. 72

Acknowledgements ............................................................................................... 80

Figures ................................................................................................................... 82

Tables .................................................................................................................... 98

Page 13: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

TABLE OF CONTENTS (Continued)

Page

Chapter 4 – An exception to the rule? Poor health of the high intertidal seaweed Fucus

distichus in the low intertidal and a test of possible mechanisms.................. 104

Abstract ............................................................................................................... 104

Introduction ......................................................................................................... 105

Methods ............................................................................................................... 111

Results ................................................................................................................. 119

Discussion ........................................................................................................... 123

Acknowledgements ............................................................................................. 131

Figures ................................................................................................................. 133

Tables .................................................................................................................. 142

Chapter 5 – General conclusion ................................................................................. 144

Bibliography .............................................................................................................. 150

Appendices ................................................................................................................. 161

Appendix A – Chapter 2 Supplemental Figures & Tables ................................. 162

Appendix B – Chapter 4 Supplemental Figures & Tables .................................. 166

Page 14: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

LIST OF FIGURES

Figure Page

2.1. Map of sampling locations. ................................................................................... 37

2.2. Mean (± SE) elemental content and ratios, by species. ........................................ 38

2.3. The slope of the relationship between N-availability and C:N across each

quantile of N-availability. ............................................................................................ 39

2.4. Regional differences in mean (± SE) C:N. ........................................................... 40

2.5. Differences in mean (± SE) C:N by cape. ............................................................. 41

2.6. Differences in mean (± SE) C:N by site. .............................................................. 42

2.7. Mean (± SE) macrophyte C:N for each month. .................................................... 43

2.8. Interannual variability in mean (± SE) macrophyte C:N for all species

combined. ..................................................................................................................... 44

3.1. Map of sampling locations. ................................................................................... 82

3.2. Mean growth rate (mm/day) of S. sessilis across months, shown by year. ........... 83

3.3. Growth rates vary among regions for S. sessilis. .................................................. 84

3.4. Variation in growth rates across capes for S. sessilis............................................ 85

3.5. S. sessilis growth rate across sites, from north to south. ....................................... 86

3.6. Mean growth rate (mm/day) of P. scouleri across months, shown by year.......... 87

3.7. Regional variation in growth rates for P. scouleri. ............................................... 88

3.8. Variation in growth rates among capes for P. scouleri. ........................................ 89

3.9. P. scouleri growth across sites. ............................................................................. 90

3.10. Nutrient availability ([NO3-1

] uM) averages for each site over the study

period. .......................................................................................................................... 91

3.11. The slope of the relationship between N-availability and growth rate of S.

sessilis and P. scouleri. ................................................................................................ 92

Page 15: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

LIST OF FIGURES (Continued)

Figure Page

3.12. Light availability shown as the average downwelling attenuation coefficient

(Kd) for each site over the study period. ...................................................................... 93

3.13. The slope of the relationship between photosynthetically active radiation

(PAR; Kd) and growth rate of S. sessilis and P. scouleri. ............................................ 94

3.14. S. sessilis %C, %N, and C:N across growth rates. .............................................. 95

3.15. The relationship between C:N and growth in S. sessilis across all three study

years and seasons ......................................................................................................... 96

3.16. P. scouleri %C, %N, and C:N across growth rates. ............................................ 97

4.1. Photos of transplanted Fucus individuals at the end of the three-month field

experiment.................................................................................................................. 133

4.2. Growth of Fucus transplants to the high and low intertidal zones, as measured

by length (a) and new tip growth (b). ........................................................................ 134

4.3. New receptacle formation (a) and final receptacle proportion (b) in the high

and low zone Fucus transplants. ................................................................................ 135

4.4. Elemental content of Fucus transplants in each zone (a) %C, b) %N, c) C:N). . 136

4.5. Growth in length of transplants of both zones and the two treatments to test

the canopy competition hypothesis. ........................................................................... 137

4.6. Growth in length of transplants of both zones (black bars= high zone, white

bars= low zone) and the three treatments testing the hypothesis that herbivory sets

the lower limit in Fucus. ............................................................................................ 138

4.7. Mesocosm results showing the influence on Fucus of immersion time (a, c, e,

g) and light level (b, d, f, h). ...................................................................................... 140

4.8. Net photosynthesis (g O2 g-1

h-1

) across irradiances (µmol photons m-2

s-1

) for

Fucus individuals in full (filled circles) and shade (open circles) light treatments. .. 140

4.9. Elemental content (%C, %N, and C:N) of Fucus individuals from the

mesocosm experiment. ............................................................................................... 141

A2.1. Interannual variability in mean (± SE) macrophyte C:N for F. distichus. ....... 162

Page 16: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

LIST OF FIGURES (Continued)

Figure Page

A2.2. Interannual variability in mean (± SE) macrophyte C:N for M. splendens. .... 163

A2.3. Interannual variability in mean (± SE) macrophyte C:N for P. scouleri. ........ 164

A2.4. Interannual variability in mean (± SE) macrophyte C:N for S. sessilis. .......... 165

B4.1. Occurrence of Fucus recruits (filled circles, #/0.25m2) and primary cover of

barnacles (open circles, % cover in 0.25m2) across vertical transects running from

+0.15m to +1.1m above MLLW. ............................................................................... 166

Page 17: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

LIST OF TABLES

Table Page

2.1. The hierarchical structure of study sites nested within capes, which are within

three regions. ................................................................................................................ 45

2.2. Study species and details on sample collection. ................................................... 46

2.3. Differences in log10(C:N) among regions for all four species. ............................. 47

2.4. Pairwise comparisons in log10(C:N) among capes and within-cape site

differences for F. distichus. ......................................................................................... 48

2.5. Pairwise comparisons in log10(C:N) among capes and within-cape site

differences for M. splendens. ....................................................................................... 49

2.6. Pairwise comparisons in log10(C:N) among capes and within-cape site

differences for P. scouleri. ........................................................................................... 50

2.7. Pairwise comparisons in log10(C:N) among capes and within-cape site

differences for S. sessilis. ............................................................................................. 51

2.8. Effect of month and site on C:N in each species. ................................................. 52

2.9. Effect of year and location (region, cape, and site analyzed separately) on

C:N for all species combined. ...................................................................................... 53

2.10. Results from likelihood ratio tests to determine the most parsimonious

model fit for determining the influence of environmental drivers on C:N. ................. 54

2.11. Results from mixed effects models on the influence of environmental

drivers on C:N. ............................................................................................................. 55

2.12. The independent contribution (%I) of each environmental driver in

explaining C:N for all species combined and each species individually. .................... 56

3.1. S. sessilis pairwise comparisons among months. .................................................. 98

3.2. Regional differences in growth rates of S. sessilis. ............................................... 99

3.3. Cape- and site-scale differences in growth rates of S. sessilis. ........................... 100

3.4. P. scouleri pairwise comparisons among months. .............................................. 101

Page 18: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

LIST OF TABLES (Continued)

Table Page

3.5. Regional differences in growth rates of P. scouleri. ........................................... 102

3.6. Cape- and site-scale differences in growth rates of P. scouleri. ......................... 103

4.1. Results from generalized linear models relating each metric of Fucus growth,

reproductive status, and elemental content to zone. .................................................. 142

4.2. Results of two-way ANOVAs testing the influence of both factors, and their

interaction, on each photosynthetic parameter: Pmax, Respiration, Slope, and

Compensating Irradiance. .......................................................................................... 143

B4.1 Results for drop-in-deviance tests comparing generalized linear models

including treatment nested within zone to models including only zone for field

experiment data. ......................................................................................................... 167

Page 19: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

The roles of nutrient dynamics, oceanography, and light in the structure and

function of rocky intertidal macrophyte assemblages

CHAPTER 1 – GENERAL INTRODUCTION

It is increasingly being recognized that bottom-up regulation of communities by

resources such as nutrients plays an important structuring role in communities. Resource

subsidies that flow across ecosystem boundaries form the basis for these bottom-up

impacts in open systems. Through the provision of nutrients in particular, these subsidies

can enhance autotroph productivity (Polis and Hurd 1996, Gorman et al. 2009), which

can then influence trophic structure and interactions (Polis et al. 1997, Marleau et al.

2010). The relationship of the primary producer community to nutrient availability is well

established in some systems, such as grasslands (Tilman 1984, 1985), but in marine

systems there remains a gap in our understanding of how resource availability affects

macrophyte communities.

Benthic nearshore habitats are ideal systems for the study of bottom-up processes.

In these habitats local ecological dynamics are coupled to oceanographic processes via

coastal upwelling, which draws deep, nutrient-rich ocean waters to the nearshore. In the

California Current Large Marine Ecosystem (CCLME) in the Northeast Pacific this

process varies spatially, and interacts with coastal geomorphology and bathymetry to

result in complex dynamics in the provision of nutrients to intertidal habitats (Graham et

al. 1992, Castelao and Barth 2005, Kirincich et al. 2005, Keister et al. 2009). This

variation relates as well to differences in the nearshore light environment driven by

phytoplankton blooms (Chan et al. 2008, Kavanaugh et al. 2009). Critically, these

Page 20: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

2

processes are expected to fundamentally change due to anthropogenic climate change

(Bakun 1990), with some of these predictions already realized in the CCLME (Iles et al.

2012). This change underscores the critical need for research into the role of the physical

environment in shaping benthic communities.

It is well established that the variation in oceanographic context throughout the

CCLME has consequences for local ecological communities (Menge et al. 2004, Menge

and Menge 2013). Top-down impacts of predators on their prey are modulated by the

strength of upwelling (Sanford 2002, Menge et al. 2003, 2004), as are herbivore-producer

interactions (Freidenburg et al. 2007). Variation in recruitment, due to upwelling and

changes in the nearshore retention of upwelled waters influenced by bathymetry, can also

have profound effects on community structure (Connolly et al. 2001). Despite this

impressive legacy of research into local and regional community dynamics in nearshore

habitats, there remain critical gaps in our knowledge, notably about the role of bottom-up

influences on benthic macrophyte assemblages (Menge 2004, Menge and Menge 2013).

Intertidal macrophyte assemblages, composed of seaweeds and surfgrasses, are

some of the most productive in the world, in large part due to the influence of upwelling

(Mann 1973, Ramirez-Garcia et al. 1998, Blanchette et al. 2009). At large spatial scales,

the structure of macrophyte communities varies in concert with regional and mesoscale

differences in oceanography (Broitman et al. 2001, Nielsen and Navarrete 2004, Wieters

et al. 2009). These macroscale patterns illustrate the potential for macrophyte

assemblages to be influenced by oceanographic subsidies such as nutrients, but the

precise ways in which that influence occurs are unknown. This issue has also received

Page 21: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

3

attention at more local scales. In a set of tidepool mesocosms, experimental nutrient

loading altered the structure and function of macrophyte assemblages, decreasing

diversity while at the same time increasing their abundance (Nielsen 2003). In other

studies on open coast macrophytes though, the impact of nutrients has been less evident

and uncertainty remains as to when and where these subsidies are important (Fujita et al.

1989, Pfister and Van Alstyne 2003, Guerry et al. 2009) The question of how these

observations integrate with the large scale patterns described above may enable us to

improve our understanding of how upwelling and nutrient availability shape macrophyte

assemblages.

My research was motivated by questions about how oceanographic subsidies and

the physical environment influence intertidal communities. The role of nutrients in

benthic ecological communities is inherently complex and interacts with other

environmental variability such as light availability, upwelling, and climatic variables. In

Chapters 2 and 3 I take a large-scale approach to these issues. Chapter 2 investigates the

coupling of regional oceanographic drivers with local intertidal macrophyte assemblages.

This investigation of drivers of variation in tissue nutrient content in a diverse group of

intertidal macrophyte species uses a dataset spanning ten years and over 900 km of

coastline in Oregon and California. By analyzing not only spatiotemporal variability but

also the influence of nutrient availability, upwelling, and climate variables, I am able to

reveal both species-specific and systematic influences on macrophyte nutrient content.

Macrophyte nutrient content, and its relationship to nutrient availability, varied among all

species. Across species though, I found consistent temporal variation in tissue nutrient

Page 22: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

4

content that was coherent with regional and basin-scale oceanographic patterns. Spatial

patterns also met my predictions that regional oceanography plays an important role in

the nutrient dynamics of intertidal macrophytes.

In Chapter 3, I further investigated two dominant low intertidal species,

Saccharina sessilis, a kelp, and Phyllospadix scouleri, a surfgrass, to determine how their

growth varies across a large biogeographic region, and how this interacts with nutrient

content. I found contrasting drivers of growth between species. Saccharina sessilis

growth was associated with both nutrient and light availability, but the relationships

varied and resource availability did not always lead to increased growth as predicted. On

the other hand, P. scouleri growth was positively associated with increasing nutrient and

light availability, but only up to intermediate levels of resource availability after which

the relationships changed. Despite these differences, and the role of nutrient availability

in modulating growth rates, I did not observe strong relationships between growth and

tissue nutrient content in these species. Nonetheless, results for both species signaled a

reliance on the process of upwelling at the regional- and meso- scales.

Chapter 4 takes a very different approach to asking how the physical environment

shapes macrophyte communities. In this study, I focused in on the local scale influences

on one species’ vertical distribution. This study was motivated by results from Chapter 2

which showed that the high intertidal seaweed, Fucus distichus, had lower nutrient

content than other species, suggesting a potential effect of tidal height on nutrient status

of this species. Using field and laboratory experiments, I tested the classical model of

rocky intertidal zonation, which predicts that lower limits are set by biotic interactions

Page 23: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

5

while upper limits are set by abiotic forces, to determine what factors set elevational

limits for F. distichus. Field reciprocal transplant experiments did not support the

hypothesis that biotic interactions set the lower limit for this species: all low zone

transplants performed worse than high zone transplants, regardless of the presence of

grazers or a macrophyte canopy. To determine the mechanism behind this result, I

performed a laboratory mesocosm experiment to test two possible physical mechanisms

that differ between zones: immersion time and light availability. I found that decreased

light availability negatively impacted F. distichus, providing further evidence for a

physical mechanism in setting the lower distribution limits in this species. In contrast to

Chapters 2 and 3, Chapter 4 illustrates an example of local influence of the physical

environment on macrophyte communities.

In my dissertation, I investigated how the complexity of environmental variation,

especially nutrient dynamics, in the CCLME influences intertidal macrophytes across

both the biogeographic scale of hundreds of kilometers down to the local scale of the

difference between zones on the order of meters. The diversity of species I investigated is

reflected in the diverse ways that macrophytes relate to environmental variation. My

findings suggest that regional variation in oceanographic drivers of nutrient availability is

manifest in both tissue nutrient content and growth of macrophytes, but the extent to

which additional factors interact with this varies. Both nutrient and light availability

played a role in driving macrophyte growth rates, and the patterns observed in growth

relate to upwelling dynamics. Finally, new insights into the role of light at both scales of

inquiry add to a small but growing body of evidence that this resource can shape

Page 24: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

6

intertidal communities. Combined, the findings discussed here contribute to our

understanding of the mechanisms through which oceanography, nutrient dynamics, and

light influence intertidal macrophyte communities.

Page 25: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

7

CHAPTER 2 – ORGANIZATION OUT OF CHAOS? SEPARATING THE

EFFECTS OF SPECIES, TEMPORAL VARIATION, CLIMATE, AND SPATIAL

SCALES AS DETERMINANTS OF INTERTIDAL MACROPHYTE NUTRIENT

DYNAMICS

ABSTRACT

Although nutrient availability is an important bottom-up subsidy in ecosystems,

its role in structuring benthic ecological communities is poorly known. For intertidal

macrophytes in coastal ecosystems, regional-scale oceanographic processes such as

upwelling largely drive nutrient availability. Hence, coastal oceanography can be an

important factor linking local areas across large spatial scales, and resolving the patterns

of rocky intertidal macrophyte nutrient content (measured as the ratio of carbon to

nitrogen, or C:N) may provide insight to the coupling of local ecological and regional

oceanographic processes. To examine biogeographic patterns of C:N ratios in intertidal

macrophytes, and their relationship to oceanographic conditions, we conducted a study of

these factors for four ecologically dominant species across 900 km of coastline in the

northern California Current Large Marine Ecosystem (CCLME) over a 10 year period.

Our research allowed us to examine the relationship of macrophyte C:N to environmental

forcing at the spatial scales appropriate to assess C:N—environment links in upwelling-

influenced habitats. We found that nutrient content of this diverse suite of rocky intertidal

macrophytes varied among species, and that C:N in all but one species was negatively

related to nitrogen availability. Patterns of spatiotemporal variability in C:N within

species were largely consistent with regional-scale predictions for decreased C:N in areas

of increased upwelling strength and persistence (i.e., decreasing latitudes), but patterns of

Page 26: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

8

meso- and local-scale variation were also evident for some species. Finally, interannual

variability was remarkably consistent across species, further implicating oceanographic-

scale factors in driving C:N.

INTRODUCTION

Understanding the influence of subsidies, such as nutrients, that flow across

ecosystem boundaries has important implications for our understanding of the drivers of

ecosystem processes such as productivity (Polis and Hurd 1996, Gravel et al. 2010,

Marleau et al. 2010). Nutrient input in particular is a major subsidy in open ecological

systems, influencing autotrophs directly by stimulating productivity, through which food

web structure is altered (Polis et al. 1997). It is well known that nutrient availability plays

an important role in determining productivity and bottom-up regulation of communities

(Dugdale 1967, Vitousek and Howarth 1991, Menge 1992), but in nearshore marine

communities these relationships are not resolved partially due to the complex

environmental variation in these systems.

In rocky intertidal habitats, subsidies influence communities via oceanographic

processes, which link areas across space and time (Menge and Menge 2013). Large

multicellular autotrophs (macrophytes) contribute greatly to the high levels of

productivity in these systems (Mann 1973, Smith 1981), provide habitat and structural

complexity (Dayton 1975, Bracken et al. 2007, Christie et al. 2009), and play important

community roles in local ecological interactions (Connell 1972, Dayton 1975, Worm and

Chapman 1996). Macrophytes rely on ecological subsidies, such as nutrients and light,

and through this reliance and their role in intertidal communities they form a link

Page 27: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

9

between benthic and pelagic habitats. Although bottom-up forces influence rocky

intertidal community regulation, their relative influence can vary on a site to site basis,

suggesting a role of nearshore oceanographic context and local ecological interactions in

driving these differences (Menge et al. 1997, Burkepile and Hay 2006, Bulleri et al.

2012). Thus, developing an understanding of how macrophytes reflect oceanographic

processes can provide insight to the role of subsidies in rocky intertidal systems such as

the California Current Large Marine Ecosystem (CCLME). To date, however, relatively

little research has been conducted at the appropriate temporal and spatial scales to

address these questions in upwelling-influenced habitats (but see for example:

Freidenburg et al. 2007, Wieters et al. 2009, Menge et al. In Prep.).

Across the CCLME the supply of nutrients is highly dynamic in space and time

(Menge et al. In Prep.). Nitrogen, the primary limiting macronutrient in marine

ecosystems, is delivered to the intertidal zones of the CCLME in the form of nitrate

through the process of upwelling, which varies seasonally. Upwelling generally starts in

March and lasts until September (Huyer 1983), but this seasonality varies with latitude

(Iles et al. 2012, Menge and Menge 2013). California experiences more persistent

upwelling and Oregon has more intermittent conditions, due to different seasonal wind

regimes and the influence of coastal topography and bathymetry. Additionally, the timing

of the spring transition to upwelling season is variable and anomalies in this timing can

have major ecosystem consequences (Peterson et al. 2006, Barth et al. 2007). Within

upwelling seasons, weather event periodicity of days to months confers additional

complexity to upwelling regimes (Iles et al. 2012). Again, the weather event scale varies

Page 28: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

10

latitudinally with more intermittent upwelling (i.e., shorter upwelling events alternating

with periods of relaxation or downwelling) in Oregon. Finally, climate scale periodicities

of the El Niño-Southern Oscillation (ENSO, 3-7 year cycles), the North Pacific Gyre

Oscillation (NPGO, 10-15 year cycles), and the Pacific Decadal Oscillation (PDO, 20-30

year cycles) operate on even longer cycles, and are associated with variability in key

ecological processes (Francis and Hare 1994, Roemmich and McGowan 1995, Mantua et

al. 1997, Chavez et al. 2003, Peterson and Schwing 2003, Menge et al. 2008, 2009).

In addition to the spatiotemporal variability of the upwelling process itself,

geographic and bathymetric features influence the way upwelled waters interact with the

shore. Major coastal headlands (capes) strengthen winds locally leading to stronger

upwelling in these locations (Graham et al. 1992, Graham and Largier 1997, Keister et al.

2009). Further, continental shelf width affects nearshore retention of upwelled waters

(Menge et al. 1997, Castelao and Barth 2005, Kirincich et al. 2005). Areas where the

continental shelf is wide, such as Heceta Bank off central Oregon (adjacent to Cape

Perpetua) tend to have longer retention time and greater onshore transport, leading to

community differences at these sites, including higher abundance of sessile invertebrates

(Menge et al. 1997) and higher incidence of nearshore phytoplankton blooms (Menge and

Menge 2013). Within- or among-site local variation can also occur, often related to

mesoscale circulation of upwelled waters and coastal topography (Legaard and Thomas

2006). Thus, oceanographic processes in the CCLME are complex and highly variable

which may be reflected in intertidal macrophyte nutrient dynamics.

Page 29: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

11

Disentangling the variation in, and drivers of, macrophyte nutrient dynamics can

be approached by measuring carbon to nitrogen (C:N) ratios. The nitrogen content of

macrophytes provides an integrated metric of their nutrient environment (Atkinson and

Smith 1983, Duarte 1992), and in temperate marine systems there is evidence that C:N of

macrophytes reflects spatiotemporal dynamics in nutrient supply (Fujita et al. 1989,

Wheeler and Björnsäter 1992, Blanchette et al. 2006).

In addition to environmental forcing, interactions with other organisms, be they

direct or indirect, can also play a role in mediating nutrient dynamics in intertidal

macrophytes (Wootton 1991, Bracken and Stachowicz 2007, Aquilino et al. 2009).

Nearshore phytoplankton blooms associated with upwelling, and their effect on inner

shelf physical and biotic factors, can increase the complexity of organism-nutrient

interactions. For example, phytoplankton may preemptively consume nitrogen in

upwelled waters before it reaches the shoreline, affecting the availability of upwelled

nitrogen to intertidal macrophytes and thereby contributing further to local- or site-scale

variation (Kavanaugh et al. 2009). Interactions can also play a role in modulating

autotroph nutrient content both negatively, through the mechanism of herbivore-induced

damage to tissues that are responsible for nutrient uptake (Bracken and Stachowicz

2007), and positively, through consumer-mediated nutrient recycling (Bracken et al.

2007, Aquilino et al. 2009).

Organismal-level processes or traits can also affect nutrient content and C:N in

macrophytes. Intertidal macrophytes are morphologically, taxonomically, and

physiologically diverse and possess variable strategies for assimilating carbon and

Page 30: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

12

nitrogen, so nutrient content can vary markedly by species. For example, the mechanisms

of nutrient acquisition and nutrient requirements are strongly linked to taxonomy. The red

algae (Phylum Rhodophyta) require a high supply of nitrogen to produce accessory

pigments like phycoerythrins, which are unique to this group, so they would be expected

to have decreased C:N. Seagrasses have less variable and higher carbon content

compared to macroalgae, which is associated with their greater requirement for structural

carbon (Duarte 1990). Some species or groups, such as the kelps and fucoids (Topinka

and Robbins 1976, Sjøtun et al. 1996), have the capacity to take up excess nutrients when

supply is high, allowing them to decouple nutrient content from nutrient availability.

Further, seaweeds only assimilate nutrients when immersed (Lobban and Harrison 1994),

and uptake rates have been shown to vary across tidal heights (Phillips and Hurd 2003,

Bracken et al. 2011), so a species’ vertical distribution can influence nutrient content.

Thus, through these organismal-level processes, in addition to environmental variation

described above, macrophyte nutrient dynamics can vary both among and within species.

The goal of this study was to deepen and expand our understanding of the

influence of oceanic subsidies on rocky intertidal benthic macrophytes. We investigated

this relationship using macrophyte tissue nutrient content. We conducted our study at a

biogeographic scale to encompass the temporal and spatial scales at which oceanography

and basin-scale processes vary in this system, enabling us to address our study questions

at the scales appropriate for upwelling-influenced Large Marine Ecosystems. In

particular, we investigated four questions, which we addressed through six hypotheses:

Q1: What are the differences in C:N among species?

Page 31: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

13

H1: Species will vary predictably in C:N ratios, based on (e.g.) tidal height,

presence of accessory pigments, and capacity for luxury storage.

Q2: What is the relationship between water column nitrogen availability (N-

availability) and macrophyte C:N in rocky intertidal habitats?

H2: N-availability will be negatively related to C:N in all species; that is, with

higher water column N-availability, tissue nitrogen content will be higher, resulting in

lower C:N ratios.

Q3: What are the spatial and temporal dynamics of macrophyte C:N?

H3.1: Spatially, patterns of C:N will correspond with regional-scale differences in

upwelling, with sites with stronger and/or more persistent upwelling (California) in

general having lower C:N. That is, C:N will decrease with decreasing latitude. Cape- and

site-scale differences will reflect regional differences.

H3.2: C:N will vary by year. Interannual patterns of C:N will cohere with

interannual oceanographic patterns in the CCLME (see H4).

H3.3: Seasonal patterns in C:N will reflect the seasonal nature of upwelling, with

lower nutrient enrichment evident in macrophyte samples in the spring and higher

nutrient content in the summer months.

Q4: How do different environmental drivers contribute to, or help explain,

variation in C:N?

H4: Nutrient availability and upwelling will have a greater contribution to C:N

compared to basin-scale processes (e.g. ENSO, NPGO). During periods of stronger than

usual upwelling, C:N will be low, and vice versa. C:N will also bear a signature of

Page 32: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

14

fluctuations in the El Niño-Southern Oscillation, being lower during the cold phase (La

Niña). Further, fluctuations in the North Pacific Gyre Oscillation (NPGO) will influence

macrophyte C:N negatively, with higher values of the NPGO (indicating stronger

currents) associated with lower values of C:N. As these environmental drivers are not

independent of one another, we expect to observe interactions among variables.

METHODS

Study sites are located along an approximately 900 km stretch of the Oregon and

Northern California coast in the northern CCLME (Fig. 2.1). This study used a spatially-

nested study design to capture the three major scales of environmental variation in this

system (region-, cape-, and site- scale). Sites were nested within large coastal headlands,

or capes, which were then grouped within regions; these are elaborated here from north to

south (Table 2.1). In northern Oregon, Cape Foulweather (CF) includes the sites Fogarty

Creek (FC), Boiler Bay (BB), and Manipulation Bay (MB), and Cape Perpetua (CP)

includes Yachats Beach (YB), Strawberry Hill (SH), and Tokatee Klootchman (TK). In

southern Oregon, Cape Arago (CA) is north of Cape Blanco (CB), which includes sites

CB North (CBN), CB South (CBS), Port Orford Head (POH), and Rocky Point (RP).

Finally, our sampling in northern California included Cape Mendocino (CME), which

consists of CME North (CMEN), CME South (CMES), Kibesillah Hill (KH), and

MacKerricher State Park (MSP). Point Arena (PTA) was the southernmost cape,

consisting of Moat Creek (MC) and Bodega Marine Lab (BML).

At each site, we sampled four intertidal macrophyte species selected because of

their high abundance, presence along the full study range, and representation of the

Page 33: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

15

diversity of macrophytes along this range: the high intertidal brown alga Fucus distichus

subsp. evanescens (C. Agardh) H.T. Powell, the low to mid intertidal red alga Mazzaella

splendens (Setchell & N.L. Gardner) Fredericq, the low intertidal surfgrass Phyllospadix

scouleri (W.J. Hooker), and the low intertidal brown kelp Saccharina sessilis (C. Agardh)

Kuntze. We collected samples biweekly (for all species except F. distichus which was

collected monthly) from March through September in 2000-2004, 2006, and 2008-2010,

although P. scouleri was only collected for 2006, 2008-2010 (Table 2.2).

Collection methods varied among species in order to account for different growth

habits of each species. S. sessilis was sampled by collecting approximately one-inch

square sections of blades from twenty-one separate individuals. Similarly, F. distichus

samples were collected from twenty-one separate individuals that were then sorted into

three replicates (of seven tissue samples each) for processing and analysis. For M.

splendens and P. scouleri, one handful of plant tissue from each of three distinct patches

was collected and later separated into three equal replicates for processing and analysis.

We placed all samples in plastic zip-top bags in the field and subsequently stored them at

-20°C. To prepare samples for C:N analysis, samples were thawed at room temperature

and all epiphytes and dead or unhealthy portions of tissue were removed. We rinsed each

sample carefully with nanopure water, then dried them in ashed foil packets at 60°C for

48 hours and stored in an airtight container with desiccant. Samples were again dried

overnight before grinding to eliminate residual moisture then ground on a Spex Sigma

Prep 8000D Mixer/Mill, and stored in ashed glass vials. C and N content (on a percent

dry weight basis) were analyzed by the Marine Science Institute at the University of

Page 34: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

16

California at Santa Barbara and at the University of Georgia Analytical Chemistry lab

with Micro-Dumas combustion on a Carlo Erba NA1500 C/H/N Analyzer.

Environmental Parameters

Ambient nutrient concentrations were obtained from samples of seawater taken

from each site. Samples were collected in triplicate from a depth of approximately 1m at

the most wave-exposed area of a site, at or as close to the time of low tide as possible.

We sampled seawater monthly from March through September on the same low tide each

month providing us with a synoptic record of environmental conditions for all sites.

Dissolved inorganic nitrate was analyzed using the cadmium reduction method (Jones

1984). For more detail see Menge et al. (In Prep.).

Upwelling and climate indices

The Bakun upwelling index provides a measure, based on wind stress, of the

strength of upwelling measured as offshore transport in m-3

s-1

100m-1

of coastline.

Positive values indicate upwelling and negative values indicate downwelling (Bakun

1967). The Bakun index was calculated using the National Oceanic and Atmospheric

Administration (NOAA) Pacific Fisheries Environmental Laboratory method accounting

for coastal angle (http://pfel.noaa.gov/) (Schwing et al. 1996). Using the coastal angle at

each site, daily upwelling indices were calculated using a radius of 0.2 degrees around

each site. All values were spatially averaged within that 0.2 degree radius of each site,

and a linear interpolation was used when data were not available. The measure of recent

upwelling used in analyses was calculated as a sum of the daily Bakun index in a 17-day

Page 35: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

17

moving window preceding each sampling date at each site. Sensitivity analysis was

performed to determine the influence of the window size by varying the window size

from 1 to 30 days and re-running the final model with each window size. The model with

a window width of 17 days had the lowest AICc score, so that window was used for all

analyses.

The multivariate El Niño-Southern Oscillation (ENSO) index (MEI) was obtained

from the NOAA Earth System Research Laboratory (see:

http://www.esrl.noaa.gov/psd/enso/mei/). The MEI is based on six environmental and

climatological variables in the tropical Pacific including pressure, temperature (air and

sea surface), wind, and cloudiness (US Department of Commerce n.d.). Negative values

of MEI indicate the ENSO cold phase (La Niña) while positive values represent the warm

phase (El Niño).

The North Pacific Gyre Oscillation (NPGO) was obtained from

http://www.o3d.org/npgo/data/NPGO.txt. The NPGO is based on observed variability in

sea surface height, and has been found to be related to salinity, nutrients, and chlorophyll

a in the Northeast Pacific and specifically in the CCLME (Di Lorenzo et al. 2008).

Statistical Analyses

All analyses were performed in R version 3.0.0 (R Core Team 2013). In all cases,

assumptions of normality and equal variances were assessed visually and C:N was log10

transformed to better meet assumptions of normality. To obtain all pairwise comparisons

we used Tukey’s HSD with the R package multcomp (Hothorn et al. 2008).

Page 36: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

18

To test the hypothesis that species would vary in their C:N (Q1, H1) we employed

mixed effects models using the R package lme4 (Bates et al. 2013). For these tests, the

only fixed effect was species. To account for the influence of spatial and temporal

variability in the difference in C:N among species, we included two random intercept

terms: one including the nested structure of sites, nested within capes, nested within

regions, and the other including the random effect of year. Because we are interested in

species-specific patterns in C:N, we analyzed each species separately.

To test the hypothesis that C:N is related to nutrient availability (Q2, H2), we took

a quantile regression approach using the R package quantreg (Koenker 2013). This

approach is appropriate for studies of limiting processes as it analyzes the relationship at

different quantiles of the explanatory variable, allowing identification of different

relationships between the response and explanatory variables across the range of the

explanatory variable (Cade et al. 1999). For each species, we analyzed the relationship

between log10C:N and N-availability (monthly averages) at twenty different quantiles (τ)

of N-availability (from 0.05 to 0.95 by increments of 0.05). We then plotted the slope (β1)

of the linear fit (y = β0 + β1x) across all quantiles (τ) to visualize how the relationship

between C:N and N-availability changes across levels of N-availability.

To investigate the spatial and temporal dynamics of C:N in intertidal macrophytes

(Q3, H3.1,3.2,3.3), we also used mixed effects models. Spatial variation in C:N was assessed

with a model including site as a fixed effect and year as a random intercept. To assist

with interpretation of pairwise comparisons, we also ran pairwise comparisons on a

model including cape and one including region. Similarly, temporal dynamics were

Page 37: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

19

analyzed using mixed effects models with month and site as multiplicative fixed effects

and cape nested within region as a random intercept. Patterns among years were analyzed

on all species combined with the interaction between year and location (region, cape, and

site were analyzed separately) as fixed effects and species as a random intercept.

To further investigate the contributions of other environmental drivers to

variability in C:N (Q4, H4), we again used mixed effects models as well as hierarchical

partitioning analysis. For the mixed effects model selection procedure, we started with a

full model with log10C:N as a function of species and environmental factors (monthly

mean nitrogen concentration [hereafter: N-availability], recent upwelling, MEI, and

NPGO) as multiplicative fixed effects, year as a random slope, and the nested spatial

structure (region/cape/site) as a random intercept. We used AICc and Likelihood Ratio

Tests to compare model fits. Once we determined the most parsimonious model, we also

compared the model fit with and without a correlation structure, testing multiple

correlation structures (first-order autoregressive, continuous autoregressive, exponential,

Gaussian, and spherical). To investigate patterns within each species, we performed

further model selection and analyses on each species individually using the same method.

To obtain estimates of the independent contribution of each of the explanatory

variables to C:N, we performed hierarchical partitioning analysis on the data first with all

species combined, and then for each species individually (MacNally 1996). For this,

log10(C:N) was the response variable, and we included environmental drivers (N-

availability, upwelling, MEI, and NPGO) and species (when applicable) as explanatory

factors. We used a Monte Carlo randomization based on 1,000 permutations of the

Page 38: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

20

hierarchical partitioning procedure to determine the statistical significance (alpha= 0.05)

of the independent contribution of each variable.

RESULTS

Differences in C:N among species

C:N differed among species (χ2

(3)= 3677.5, P< 0.0001). In pairwise comparisons,

all species differed from each other (Fig. 2.2c; P< 0.0001 for all pairwise comparisons;

M. splendens < S. sessilis < P. scouleri < F. distichus). Carbon and nitrogen content also

varied among all species (%C: χ2

(3)= 5100.3, P< 0.0001; %N: χ2

(3)= 1267.7, P< 0.0001).

For carbon, each species differed from all others (Fig. 2.2a; P< 0.0001 for all; M.

splendens < S. sessilis < F. distichus < P. scouleri). Similarly, nitrogen content of each

species was unique from all others (Fig. 2.2b; P< 0.0001 for all; F. distichus < M.

splendens < S. sessilis < P. scouleri).

Relationship between C:N and nitrogen availability

Quantile regression revealed that C:N was negatively related with N-availability

for all species except F. distichus (Fig. 2.3). That is, the higher the N-availability, the

lower the C:N ratio, or the more enriched was the algal species with nitrogen. For M.

splendens, P. scouleri, and S. sessilis, patterns were similar with negative slopes (β1)

across most quantiles of N-availability. The 0.05th

quantile of N-availability was

negatively related to C:N in P. scouleri only (no other species had a relationship at that

quantile), and the 0.95th

quantile was not related to C:N for any species. Across quantiles,

the slope of the relationship between N-availability and C:N did not vary, although it was

slightly more negative in S. sessilis, and slightly more positive in P. scouleri. The

Page 39: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

21

relationship between C:N and N-availability was not significant at any quantile for F.

distichus. Using coarser quantile bins (0.10) did not change this conclusion (data not

shown).

Spatial Variation in C:N

Pairwise differences among regions (Northern Oregon, Southern Oregon, and

Northern California) revealed regional differences in C:N, which varied among species

(Fig. 2.4, Table 2.3). F. distichus, which was only sampled in Oregon, had higher C:N in

Northern Oregon compared to Southern Oregon (Fig. 2.4a, P-adj.< 0.0001). M. splendens

did not exhibit any differences in C:N among regions (Fig. 2.4b, P-adj = 0.174 or higher

for all comparisons). P. scouleri did not vary between the Oregon regions, but C:N in

both Oregon regions was higher than in California (Fig. 2.4c, P-adj < 0.001 for both

comparisons). S. sessilis followed the opposite pattern, with no differences between

Oregon and California sites, but higher C:N in Northern Oregon compared to Southern

Oregon (Fig. 2.4d, S-Ore vs. N-Ore: P-adj.<0.0001).

Cape-wise differences in C:N largely mirrored regional differences (Fig. 2.5,

Tables 2.4-2.7). For F. distichus, capes in Northern Oregon (CF and CP) had greater C:N

than capes in Southern Oregon (CA and CB; P-adj = 0.001 or less), but there were no

within-region differences among capes (P-adj = 0.86 or more)(Fig. 2.5a, Table 2.4).

Despite its lack of regional differences, C:N in M. splendens was lower at CA (P-adj =

0.01 or lower) than all other capes except PTA (P-adj. = 0.08) (Fig. 2.5b, Table 2.5). No

other cape-wise differences were evident in M. splendens. Cape-wise differences in P.

scouleri were related to two capes, primarily, CP and PTA (Fig. 2.5c, Table 2.6). C:N at

Page 40: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

22

both of these capes differed from all others (although CP vs. CB was only marginally

significant, P-adj.= 0.052; in all others P-adj. = 0.02 or lower). C:N was lower at PTA

and higher at CP compared to all other capes. Additionally, C:N was lower at CME

compared to CB (P-adj.= 0.003). A similar pattern, with one cape standing out from all

others, was evident for S. sessilis, in which C:N at CA was lower than C:N for all other

capes (P-adj. = 0.04 or lower, Fig. 2.5d, Table 2.7). Additionally, CP had higher C:N than

CB(P-adj.<0.001).

Site-level variation in C:N was evident, to different extents, in all species (Fig.

2.6, Tables 2.4-2.7). Of the site-level differences in C:N, here we report only within-cape

differences for all capes except CA (there was no within-cape site replication at CA). F.

distichus did not have any within-cape differences among sites (Fig. 2.6a, Table 2.4). For

M. splendens, C:N at CBN was lower than C:N at all other sites at CB (Fig. 2.6b, Table

2.5; CBN < CBS = POH = RP), and within CP, YB had lower C:N than TK (YB < TK,

P-adj.< 0.01). P. scouleri had many within-cape differences among sites (Fig. 2.6c, Table

2.6). At CF, MB had higher C:N compared to FC (P-adj.< 0.01). At CB, CBN and CBS

both had lower C:N than POH and RP, but neither pair differed from each other (CBN =

CBS < POH = RP;). At CME, KH had lower C:N than all other sites at that cape (KH<

CMEN = CMES = MSP). S. sessilis had few within-cape differences (Fig. 2.6d, Table

2.7). CBS had greater C:N than CBN (CBS > CBN; P-adj.< 0.01), and YB and SH had

lower C:N than TK (YB = SH < TK).

Temporal variation in C:N

Page 41: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

23

Variations in C:N across months were similar across species (Table 2.8, Fig. 2.7).

For all species, C:N depended on the interactive effect of month and site (P< 0.001 for

all). The patterns in C:N across months varied among species. From April through June,

F. distichus did not vary in C:N, but those four months all had greater C:N than July,

August, and September (P-adj = 0.04 or less for all). Within the late summer months

(July-Sept), C:N was lower in August than July or September (P-adj. = 0.01 or less vs. P-

adj.= 1). C:N in F. distichus in March did not differ from any other months (P-adj.> 0.7

for all).

M. splendens had a similar break in C:N between seasons; however, for this

species March through May grouped together and C:N for each of those months was less

than C:N for June through August (P-adj. = 0.001 or less in all cases). September had

lower C:N than June through August (P-adj. = 0.001 or less), and did not differ from the

spring months. P. scouleri was not sampled in September, and had little seasonal

variation in C:N with March having greater values than all other months (P-adj. = 0.0002

or less) but with no other differences among months. Likewise, C:N in S. sessilis did not

show much evidence of seasonal variation either. May and August both had greater C:N

than June or July (P-adj. = 0.005 or less), but there were no other differences in C:N

among months.

Interannual variability in C:N was remarkably consistent across species and

latitude, so we combined all data to obtain an annual average C:N across all species (Fig.

2.8). For all species combined, year and location (either site, cape, or region) were related

to C:N, and the relationship with year depended on location (Table 2.9).

Page 42: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

24

Drivers of C:N

To investigate the environmental drivers of C:N in macrophytes, we first

determined the most parsimonious models. For the all-species model, the most

parsimonious model indicated by Likelihood ratio tests was a linear mixed effects model

with log10(C:N) as a function of the nested spatial structure as a random intercept

(region/cape/site), species and environmental drivers (N-availability, recent upwelling,

MEI, and NPGO) as multiplicative fixed effects, and an autoregressive correlation

structure (AR1) (L-ratio= 299.42, P< 0.0001). The model selection procedure did not

result in dropping any terms from the full model. We found that using a correlation

structure was better than omitting one (L-ratio= 138.04, p< 0.0001). However, three

correlation structures tied in AICc comparisons (first-order autoregressive, continuous

autoregressive, and exponential), so the first-order autoregressive correlation structure

(AR1) was chosen for its simplicity (Zuur et al. 2009).

For analysis of each species separately, we found that the best model was the

same for all species except S. sessilis (Table 2.10). For F. distichus, M. splendens, and P.

scouleri, the final model included log10(C:N) as the response variable, nested spatial

structure as the random intercept (region/cape/site), environmental drivers (N-

availability, recent upwelling, MEI, and NPGO) as multiplicative fixed effects up to and

including all three-way interactions, and a correlation structure. For S. sessilis, the final

model was the full model using log10(C:N) as the response variable, nested spatial

structure as the random intercept (region/cape/site), environmental drivers (N-

availability, recent upwelling, MEI, and NPGO) as multiplicative fixed effects including

Page 43: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

25

all interactions, and a correlation structure. For all species we used a first-order

autoregressive correlation structure for consistency with the all-species model.

As expected, mixed effects models revealed complex higher-order interactions of

N-availability with other environmental drivers for all species but F. distichus (Table

2.11). In M. splendens, the effect of N-availability on C:N depended on recent upwelling

and MEI (t1379= 2.32, p= 0.02). In P. scouleri, N-availability, MEI, and NPGO all

impacted C:N (three-way interaction; t443= -2.00, p= 0.04). In S. sessilis, the effect of N-

availability on C:N depended on all other factors: recent upwelling, MEI, and NPGO

(four way interaction; t1290= -3.27, p= 0.001).

For F. distichus, the relationship between C:N and MEI depended on NPGO

(t673= -2.59, p= 0.01). For each unit increase in MEI, C:N was higher by a factor of 1.14

(95% CI: 1.03, 1.26), but at different levels of NPGO, the effect of MEI on C:N

decreased by a factor of 1.0 (95% CI: 0.9999872, 1.00001).

Species identity accounted for 97.2% of the explained variation in C:N in

hierarchical partitioning analysis in the all-species model (R2 = 0.53, Table 2.12).

Although the combined contributions of N-availability, recent upwelling, MEI, and

NPGO together explained a small fraction of the variability in single species models

(15% of the variability in C:N to less than 6%) independent contributions of these

environmental drivers to C:N were significant, with some differences among species. In

all models, nitrogen availability helped explain C:N variability. For example, for M.

splendens and P. scouleri, N-availability was the most important environmental driver

Page 44: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

26

with an independent contribution to the explained variation of 78.6% and 53.5%,

respectively (R2= 0.014 and R

2= 0.141, respectively).

Upwelling was also an important independent contributor. For S. sessilis and F.

distichus, recent upwelling had the highest independent contribution to explained

variation in hierarchical partitioning (66.0 and 57.4%, respectively). Additionally, in the

all species model, recent upwelling was also an independent contributor to variation in

C:N. Recent upwelling explained little variability in M. splendens.

Of the climate indices, the independent effect of MEI explained variation in C:N

only for F. distichus and M. splendens (13.6% and 10.4%, respectively). Finally, the

independent contribution of NPGO explained variation in C:N in the all-species model

and for each species except M. splendens. Of the other models, NPGO contributed only

0.28% to the explained variation in the all-species model, and 17.6%, 22.7%, and 8.4%

for F. distichus, P. scouleri, and S. sessilis, respectively.

DISCUSSION

Spatiotemporal patterns in C:N of all four species were largely consistent with our

expectations of the likely environmental drivers underlying the pattern. We expected, and

observed, a decrease in C:N with decreasing latitude due to the influence of differing

upwelling regimes in California and Oregon. Perhaps most strikingly, interannual

variability in C:N was remarkably consistent across both species and latitude and these

patterns cohere with oceanographic patterns for those years. For all four species, three

seaweeds and one seagrass, our results suggest that C:N also has an intrinsic component

that is unique to a species or group of species.

Page 45: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

27

Species differences

Our results were consistent with our hypothesis that species would differ in their

C:N (Q1, H1). The four species we investigated are taxonomically diverse, F. distichus

and S. sessilis, both brown algae, M. splendens, a red alga, and P. scouleri, a seagrass.

Aspects of this diversity that affect nutrient content include differences in light harvesting

pigments, carbohydrate storage and structural molecules, life histories, and body plans

(Lobban and Harrison 1994). Our observations of differences in C:N among species are

fairly consistent with our expectations involving these traits (Fig. 2.2). F. distichus had

the highest C:N and M. splendens had the lowest, as predicted based on their tidal heights

and nitrogen requirements. However, when looking at %C and %N across species, it

appears that low C:N in M. splendens is driven by its lower carbon content in addition to

its higher nitrogen content. This unexpected result may relate to its position low on the

shore where light availability may limit C-fixation. While P. scouleri’s C:N was

intermediate to the others, it actually had the highest %C and %N across species,

potentially reflecting both the increased structural carbon present in seagrasses and higher

nitrogen reserves. S. sessilis also had intermediate C:N, %C, and %N, consistent with our

expectations based on this species’ intertidal distribution at mid-low heights on the shore,

and its capacity to take up excess nutrients in times of high supply (Lobban and Harrison

1994).

Nitrogen Availability and C:N

To understand what environmental factors drive variability in C:N in intertidal

macrophytes, we first looked at the relationship between environmental nitrogen

Page 46: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

28

availability and C:N in macrophyte tissues (Q2, H2). As expected, for all species but F.

distichus C:N and N-availability were negatively correlated across most quantiles of N-

availability, supporting our hypothesis (H2). That is, higher N-availability was associated

with low C:N ratios. While the relationship between C:N and N-availability did not vary

much across quantiles of N-availability in M. splendens, slight variations were evident for

other species. For S. sessilis, the relationship appeared to become slightly more negative

for higher quantiles of N-availability, likely a signature of increased luxury uptake and

storage of nitrogen at periods of high supply. P. scouleri exhibited the opposite trend,

with slight increases in the slope of the relationship between N-availability and C:N

across quantiles. This suggests that for surfgrass, C:N and N-availability are most tightly

coupled during times of lower N-availability. In seagrasses, low and intermediate levels

of nutrient availability can stimulate growth, but at high concentrations nitrogen also can

cause toxicity (Larkum et al. 2006). The observed relationship between C:N and N-

availability in P. scouleri may reflect that relationship. The fact that C:N in F. distichus

was not related to N-availability may stem from the fact that F. distichus is a relatively

slow-growing high zone species, and it was only sampled monthly, so its nutrient content

may not reflect environmental processes on the relevant time scales; that is, C:N and N-

availability may be decoupled in this species. Further, while C:N was sampled biweekly

(monthly for F. distichus), N-availability values used here are monthly averages. Despite

that temporal mismatch, the patterns seen in the relationships between C:N and N-

availability are largely consistent with our expectations. Looking at how these patterns

Page 47: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

29

varied spatiotemporally, we can see the hallmarks of other environmental drivers in

addition to N-availability.

Spatiotemporal variation in C:N

The pattern of decreasing C:N from north to south, as well as the consistent

pattern in interannual variability across species and latitude, are both suggestive of

environmental forcing of C:N at the regional scale. Nutrients are delivered to intertidal

habitats in this system primarily via upwelling. Spatially, we would expect C:N to

decrease from north to south due to the presence of stronger and more persistent

upwelling, and the resulting higher delivery of nutrients, in California, and indeed this is

generally the pattern we observed for all species but M. splendens. However for all

species but F. distichus we also observed varying degrees of cape- and site-scale

differences.

At the cape- and site-scales, species followed different patterns but there were

some consistencies. Generally, M. splendens and S. sessilis both varied across cape and

site levels similarly, whereas P. scouleri followed somewhat different patterns. Within-

region differences among capes were largely consistent with regional patterns in C:N and

upwelling regimes, especially for F. distichus in which capes separated exactly by region.

For the remaining three species, one or two capes were consistently set apart from most

others. For M. splendens and S. sessilis, this was Cape Arago on the southern Oregon

coast, which is known to have an oceanographic regime more similar to California than

to Oregon capes. Cape Arago is characterized by strong upwelling and offshore transport

(Connolly et al. 2001, Sanford et al. 2003). For P. scouleri, C:N at Cape Perpetua, a

Page 48: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

30

northern Oregon cape with weaker inner shelf currents and more retentive waters, was

generally higher than other capes. At Point Arena, a strongly upwelled California cape, it

was lower than other capes. Generally, these cape-wise differences fit within our

understanding of upwelling regimes in these areas. Interestingly, although cape scales

seem to be driving these patterns, the specific cape responsible is not consistent across

species.

Environmental variability can occur in this system even between sites within a

cape. As upwelled waters move across the cape they interact with phytoplankton and are

changed as they move downstream of an upwelling center. These spatiotemporal changes

in the upwelled waters within a cape relate to observed variation. Within-cape site

differences in both M. splendens and S. sessilis were apparent at Cape Perpetua, and for

those two species plus P. scouleri at Cape Blanco. This can be interpreted in terms of the

relation of the upwelling process to bathymetry at Cape Perpetua, and coastal

geomorphology at Cape Blanco. The wide continental shelf at Cape Perpetua causes

upwelled waters to be retained nearshore, fostering the formation of dense phytoplankton

blooms (Chan et al. 2008, Menge and Menge 2013). The resulting light attenuation has

been shown to reduce S. sessilis growth (Kavanaugh et al. 2009), and preemptive

consumption of nitrogen by phytoplankton may draw down nitrogen availability and

influence temporal dynamics of nutrient content in intertidal macrophytes. Similarly,

Cape Blanco is the westernmost cape on the Oregon coast and a strong upwelling jet

flows offshore at this location (Barth et al. 2000). At Cape Blanco, the site CBN drove

most within-cape differences, having lower C:N than other sites within this cape. CBN is

Page 49: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

31

located at the upwelling center for this cape, while other Cape Blanco sites are

downstream of the upwelling jet. Thus newly upwelled waters may influence C:N more

strongly at CBN than at other Cape Blanco sites.

In contrast to spatial variability, patterns of temporal variability among species

were consistent. Seasonally, C:N was generally higher in spring months and lower in

summer months. As a seasonal process, we would expect upwelling (i.e., high rates of

nutrient delivery) to drive down C:N ratios over the course of the upwelling season, with

the lowest values occurring at the end of the summer. With the exception of F. distichus,

in which C:N increased from August to September, this was the observed pattern.

Established relationships between autotroph nutrient content and light availability could

also influence this seasonal trend (Sterner et al. 1997).

Annual variation in C:N was consistent among species and geographically,

suggesting interannual patterns in environmental drivers may play an important role in

determining C:N. Generally, C:N was higher in the years 2000, 2005, and 2010. When

comparing this to oceanographic conditions in the CCLME during those years, we see

evidence of coherence among patterns as well as of interacting environmental conditions.

In 2005, a delayed spring transition to upwelling season strongly influenced ecosystem

processes in the CCLME, leading to unusually warm surface waters, low zooplankton

biomass, reproductive failure and mortality of some seabird populations, and recruitment

failure in some fish species, among other impacts (Peterson et al. 2006, Barth et al. 2007).

Its influence can be seen here in the increased C:N as well. On top of the anomalies in

upwelling, the negative NPGO in that year also indicates decreased nutrient availability

Page 50: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

32

to the region, and was manifest in decreased productivity across the CCLME in 2005 (Di

Lorenzo et al. 2008, Bjorkstedt et al. 2011). The fact that these conditions coincided may

have led to greater C:N values in this year than would have resulted if the effects were

separated. On the other hand, though C:N was also anomalously high in 2000 and 2010,

ocean conditions were quite different than 2005. NPGO was strongly positive in both

2000 and 2010, indicating increased nutrient availability, but upwelling was generally

normal in both of these years (Durazo et al. 2001, Di Lorenzo et al. 2008, Bjorkstedt et al.

2011). In 2000, normal upwelling conditions in the northern CCLME coincided with the

third year of La Niña conditions, circumstances that should lead to decreased C:N for that

year (Durazo et al. 2001). In 2010, the lingering effects of El Niño conditions may have

led to increased C:N, but El Niño neutral conditions had returned by early in our

sampling that year and given the upwelling and NPGO conditions it seems unlikely that

the weak El Niño could override their influence. Overall, patterns in environmental

drivers of C:N across years, thus, can be parsed out for some years (i.e. 2005), but

unexplained variation in 2000 and 2010 suggests that a more complicated interaction of

environmental drivers, spatial variability, and species responses may be responsible for

some of these patterns.

Environmental drivers of C:N vary by species

The complex interaction of environmental drivers in C:N, manifest to some extent

in spatiotemporal variation, can be seen in the results from analyses of environmental

drivers of C:N. For all species, at least three out of the four factors individually

contributed to C:N, and, as expected, either N-availability or recent upwelling accounted

Page 51: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

33

for the highest contribution to explained variance in C:N. The interaction of these factors

with either one or both basin-scale oceanographic indices, MEI and NPGO, in all species

suggests that while N-availability or upwelling may be the primary driver of C:N, its

influence is modulated by larger-scale oceanographic patterns. Integrating these complex

interactions into known patterns of spatiotemporal variation for each species, we can start

to develop a framework for what drives C:N in each of these four species.

Beyond regional patterns showing lower C:N in southern Oregon, and cape-scale

patterns consistent with the regional differences, F. distichus did not display any

additional cape- or site-scale variation. This pattern suggests that C:N in this species is

primarily influenced by basin-scale oceanographic drivers that differ little across the

study area. Further, the presence of temporal variation in this species suggests that C:N

does vary in F. distichus in relation to environmental cues. F. distichus’s relatively slow

growth rates may drive this observation, as C:N would be integrated over longer time

periods and decoupled from higher frequency environmental variation (Ang 1991). The

absence of a relationship between C:N and N-availability in F. distichus supports these

conclusions. Further bolstering this claim, the only environmental drivers that influenced

C:N in F. distichus, according to mixed effects model analysis, were the basin-scale

indices of MEI and NPGO that operate over longer periodicities than other environmental

drivers we investigated.

In contrast to F. distichus, C:N in P. scouleri appears to be very sensitive to local

environmental factors, displaying many pairwise differences at both the cape and site

scales. In addition to spatial patterns that were consistent with M. splendens and S.

Page 52: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

34

sessilis, there were more differences among California sites, and more within-cape site

differences in P. scouleri than in other species. Our analyses for this species suggested

that C:N is more strongly negatively related to N-availability when N-availability is low,

possibly reflecting nitrogen toxicity at high concentrations. C:N levels in P. scouleri were

above published critical nutrient levels for this species (Duarte 1990), suggesting that

they are not strictly nitrogen limited, this relationship to N-availability, and the sensitivity

of C:N to spatiotemporal variability shows that C:N in P. scouleri is very responsive to

the physical environment. The total variation explained by the four environmental drivers

analyzed was higher in P. scouleri than any other species, and the interaction between N-

availability, MEI, and NPGO further underline the responsiveness of C:N in P. scouleri

to environmental variability.

In M. splendens, C:N was consistently negatively related to N-availability and N-

availability had the highest independent contribution to M. splendens C:N. The higher

requirement for nitrogen in the red algae, of which this species is a member, likely

influences this relationship. Spatial variability in this species was very similar to that of S.

sessilis and was related to known mesoscale modulation of the upwelling process by

bathymetry and coastal geomorphology. S. sessilis, though, had a different relationship

between C:N and N-availability, wherein the greater capacity of this species for luxury

uptake of nutrients was evident. As N-availability increased, the slope of the relationship

became more negative. Instead of the decoupling of C:N from N-availability that we

would expect to see as a signature of luxury uptake, we actually observe a stronger

relationship at high quantiles of N-availability. The observed four-way interaction in this

Page 53: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

35

species among all environmental variables may stem from this capacity for luxury uptake

as well. Overall, the similar patterns observed in M. splendens and S. sessilis may be

attributed to their need (in the case of M. splendens) or ability (in the case of S. sessilis)

to take up excess nitrogen.

Conclusion

The patterns and processes underlying variation in macrophyte C:N ratios are

wide-ranging and complex, and in this dynamic ecosystem where the environment is

constantly changing these patterns can be difficult to parse. However, by evaluating C:N

in four dominant intertidal species we have shown that there are some generalities, and

some peculiarities, of drivers of C:N in intertidal macrophytes. C:N differed among all

four species, on top of which species showed different patterns of spatial variation and

relationships to environmental drivers. Both spatial variation and environmental drivers

were consistent with regional forcing of C:N; however, for all species but F. distichus,

meso- and local-scale variation was also important. Despite these variations among

species though, patterns of interannual variation in C:N were relatively consistent across

species and were driven by the interactive effects of oceanographic and climate

processes. The influence of these processes on C:N gives us further insight into how

oceanic subsidies affect intertidal macrophytes in this dynamic system.

ACKNOWLEDGEMENTS

We wish to thank Tarik Gouhier for statistical advice and for providing the

upwelling index used in our analyses. Alison Iles, Erin Gorsich, Bree Beechler, Elizabeth

Page 54: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

36

Cerny-Chipman, and members of Scientific Writing and Ethics (Z599) in Winter 2013 all

provided insightful feedback on the manuscript. At both institutions where we conducted

this research countless volunteers, interns, technicians, and graduate students assisted

with sample collection and processing. Of particular note, at Oregon State University,

Erin Richmond, Mae Noble, Gayle Murphy, Camryn Pennington, Ryan Craig, Megan

Poole, Lindsay Hunter, Angela Johnson, Jonathan Robinson, Shawn Gerrity, Jeremy

Henderson, and Mary Ellis were all instrumental in collecting and/or processing samples.

At Sonoma State University, Adele Paquin, Russel DiFiori, Beatrice Cortina, Terra

Hazen, Rachel Francis, Samuel Briggs, Kandis Gilmore and Trevor Mangor all spent

considerable time and effort collecting and processing samples. We also thank Susan

Williams (UC Davis, Bodega Marine Laboratory) for generously allowing the use of her

algal grinder. This research was supported by a National Science Foundation Graduate

Research Fellowship to SLC, NSF grant numbers OCE 0726983 and 1061233 to BAM,

FC, KJN, and SDH, NSF grant number DEB 1050694 to BAM, endowment funds from

the Wayne and Gladys Valley Foundation, grants from the A. W. Mellon Foundation to

BAM and Jane Lubchenco, and the David and Lucile Packard and Gordon and Betty

Moore Foundations for support of the Partnership for Interdisciplinary Studies of Coastal

Oceans (PISCO) consortium BAM, JL, lead PIs.

Page 55: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

37

FIGURES

Figure 2.1. Map of sampling locations. Each closed circle indicates one sampling site,

although some points overlap. Sites are listed from north to south within a cape, and

groupings by cape are as follows (from north to south): Cape Foulweather (sites: Fogarty

Creek (FC), Boiler Bay (BB), Manipulation Bay (MB)); Cape Perpetua (sites: Yachats

Beach (YB), Strawberry Hill (SH), Tokatee Klootchman (TK)); Cape Arago (CA; no

replicate sites); Cape Blanco (sites: Cape Blanco North (CBN), Cape Blanco South

(CBS), Port Orford Head (POH), Rocky Point (RP)); Cape Mendocino (sites: Cape

Mendocino North (CMEN), Cape Mendocino South (CMES), Kibesillah Hill (KH),

MacKerricher State Park (MSP)); and Point Arena (sites: Moat Creek (MC), Bodega

Marine Lab (BML)).

Page 56: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

38

Figure 2.2. Mean (± SE) elemental content and ratios, by species (a. % C, b. % N, c.

C:N). Letters indicate differences among species (Tukey’s HSD, alpha= 0.05).

Page 57: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

39

Figure 2.3. The slope of the relationship between N-availability and C:N across each

quantile of N-availability from 0.05 to 0.95 by increments of 0.05. Species are shown on

separate plots (a: F. distichus, b: M. splendens, c: P. scouleri, d: S. sessilis). Filled circles

represent a significant relationship between N-availability and macrophyte C:N at the

alpha= 0.05 level, while open circles represent non-significance.

Page 58: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

40

Figure 2.4. Regional differences in mean (± SE) C:N from north to south (left to right:

Northern Oregon, Southern Oregon, and Northern California) for each species (a: F.

distichus, b: M. splendens, c: P. scouleri, d: S. sessilis). Letters indicate differences

among regions within each species (Tukey’s HSD, alpha= 0.05). In plot a, “nd” indicates

that F. distichus was not sampled in Northern California. Note that the scales of y-axes

are different.

Page 59: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

41

Figure 2.5. Differences in mean (± SE) C:N by cape from north to south (left to right, see

Figure 1 for cape abbreviations) for each species (a: F. distichus, b: M. splendens, c: P.

scouleri, d: S. sessilis). Letters indicate differences among capes within each species

(Tukey’s HSD, alpha= 0.05). In plot a, “nd” indicates that F. distichus was not sampled

at CME or PTA. Note that the scales of y-axes are different.

Page 60: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

42

Figure 2.6. Differences in mean (± SE) C:N by site from north to south (left to right, see

Figure 1 for site abbreviations) for each species (a: F. distichus, b: M. splendens, c: P.

scouleri, d: S. sessilis). In plots a and d, “nd” indicates that F. distichus and S. sessilis

were not sampled at all sites. Note that the scales of y-axes are different.

Page 61: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

43

Figure 2.7. Mean (± SE) macrophyte C:N for each month of sampling from March to

September, averaged across years. Species are separated by row (a-b: F. distichus, c-e: M.

splendens, f-h: P. scouleri, i-k: S. sessilis). Regions are delineated by columns (left:

Northern OR, center: Southern OR, right: Northern CA), and capes are shown in different

symbols within each region (Closed circles: Cape Foulweather, Cape Arago, and Cape

Mendocino; Open circles: Cape Perpetua, Cape Blanco, and Point Arena).

Page 62: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

44

Figure 2.8. Interannual variability in mean (± SE) macrophyte C:N for all species

combined across regions from north to south (a: Northern Oregon, b: Southern Oregon, c:

Northern California). Capes within each region are shown in filled and open circles.

Page 63: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

45

TABLES

Table 2.1. The hierarchical structure of study sites nested within capes, which are within

three regions.

Region Cape Site Latitude (°N) Longitude (°W)

Fogarty Creek (FC) 44.838 -124.060

Boiler Bay (BB) 44.832 -124.059

Manipulation Bay (MB) 44.830 -124.063

Yachats Beach (YB) 44.319 -124.109

Strawberry Hill (SH) 44.250 -124.114

Tokatee Klootchman (TK) 44.207 -124.117

Cape Arago (CA) Cape Arago (CA) 43.308 -124.400

Cape Blanco North (CBN) 42.840 -124.566

Cape Blanco South (CBS) 42.834 -124.565

Port Orford Head (POH) 42.733 -124.510

Rocky Point (RP) 42.720 -124.469

Cape Mendocino North (CMEN) 40.440 -124.408

Cape Mendocino South (CMES) 40.341 -124.361

Kibesillah Hill (KH) 39.605 -123.790

MacKerricher State Park (MSP) 39.477 -123.804

Moat Creek (MC) 38.881 -123.676

Bodega Marine Lab (BML) 38.319 -123.074

Northern California

Cape Mendocino (CME)

Point Arena (PTA)

Northern Oregon

Cape Foulweather (CF)

Cape Perpetua (CP)

Southern Oregon

Cape Blanco (CB)

Page 64: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

46

Table 2.2. Study species and details on sample collection.

Species Years Collected Sites Collected (N to S) Collection Method

Fucus distichus 2000-2006, 2008-2010 FC, BB, SH, TK, CA, CBN21- 1” blade sections taken from 21

individuals, divided into 3 replicates in the lab

Mazzaella splendens 2000-2006, 2008-2010

FC, BB, MB, YB, SH, TK, CA, CBN,

CBS, POH, RP, CMEN, CMES, KH,

MSP, MC, BML

1 handful of tissue from each of 3 distinct

patches, divided into 3 replicates in the lab

Phyllospadix scouleri 2008-2010 Same as M. splendens1 handful of tissue from each of 3 distinct

patches, divided into 3 replicates in the lab

Saccharina sessilis 2000-2010

FC, BB, MB, YB, SH, TK, CA, CBN,

CBS, POH, RP, CMEN, CMES, KH,

MSP, BML

21- 1” blade sections taken from 21

individuals, divided into 3 replicates in the lab

Page 65: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

47

Table 2.3. Differences in log10(C:N) among regions for all four species. Pairwise comparisons were performed using Tukey’s

HSD, alpha= 0.05. .

Species Region Estimate Standard Error z-value Adjusted P-value

Northern OR vs. Southern OR -0.036 0.006 -6.242 <0.0001

Southern OR 0.005 0.003 -1.777 0.174

Northern CA -0.003 0.004 0.798 0.701

Southern OR vs. Northern CA -0.002 0.004 -0.497 0.871

Southern OR 0.008 0.004 -1.991 0.114

Northern CA 0.050 0.004 12.322 <0.001

Southern OR vs. Northern CA 0.042 0.004 9.660 <0.001

Southern OR 0.020 0.003 -7.390 <0.0001

Northern CA 0.010 0.004 2.322 0.051

Southern OR vs. Northern CA -0.010 0.004 -2.242 0.061

F. distichus

M. splendens

P. scouleri

S. sessilis

Northern OR vs.

Northern OR vs.

Northern OR vs.

Page 66: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

48

Table 2.4. Pairwise comparisons in log10(C:N) among capes and within-cape site differences for F. distichus. Pairwise

comparisons were performed using Tukey’s HSD, alpha= 0.05. .

Estimate Standard Error z-value Adjusted P-value Estimate Standard Error z-value Adjusted P-value

CP 0.006 0.008 -0.780 0.863 FC vs. BB 0.021 0.014 1.517 0.633

CA 0.038 0.008 4.546 < 0.0001

CB 0.040 0.008 5.042 < 0.0001

CA 0.032 0.009 3.779 0.001 SH vs. TK -0.003 0.017 0.166 1.000

CB 0.034 0.008 4.204 < 0.0001

CA vs. CB 0.002 0.009 -0.173 0.998 CA vs. CBN 0.002 0.009 -0.170 1.000

CP vs.

SiteCape

CF vs.

Page 67: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

49

Table 2.5. Pairwise comparisons in log10(C:N) among capes and within-cape site differences for M. splendens. Pairwise

comparisons were performed using Tukey’s HSD, alpha= 0.05.

Estimate Standard Error z-value Adjusted P-value Estimate Standard Error z-value Adjusted P-value

CP 0.001 0.004 -0.269 1.000 FC vs. BB -0.004 0.007 -0.534 1.000

CA 0.023 0.005 4.404 <0.001 MB -0.003 0.009 0.404 1.000

CB -0.001 0.004 -0.258 1.000 BB vs. MB 0.000 0.008 -0.010 1.000

CME 0.001 0.005 -0.236 1.000

PTA 0.004 0.006 -0.663 0.985

CA 0.021 0.005 4.154 <0.001 YB vs. SH -0.024 0.008 -2.982 0.177

CB -0.002 0.004 -0.524 0.995 TK -0.040 0.009 -4.449 <0.01

CME 0.000 0.005 0.037 1.000 SH vs. TK -0.016 0.007 2.245 0.673

PTA 0.003 0.006 -0.479 0.997

CB -0.024 0.005 4.572 <0.001

CME -0.021 0.006 3.285 0.012

PTA -0.019 0.007 2.675 0.076

CME 0.002 0.005 -0.433 0.998 CBN vs. CBS -0.047 0.008 5.816 <0.01

PTA 0.005 0.006 -0.848 0.956 POH -0.029 0.008 3.496 0.040

RP -0.032 0.008 4.157 <0.01

CBS vs. POH 0.018 0.010 -1.846 0.907

RP 0.015 0.009 -1.569 0.977

POH vs. RP -0.003 0.010 0.354 1.000

CME vs. PTA 0.003 0.007 -0.401 0.999 CMEN vs. CMES -0.022 0.012 1.783 0.929

KH 0.010 0.011 -0.906 1.000

MSP -0.015 0.014 1.052 1.000

CMES vs. KH 0.032 0.012 -2.763 0.293

MSP 0.007 0.015 -0.475 1.000

KH vs. MSP -0.025 0.014 1.840 0.909

PTA MC vs. BML 0.026 0.010 2.618 0.390

CF vs.

CP vs.

CA vs.

CB vs.

SiteCape

Page 68: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

50

Table 2.6. Pairwise comparisons in log10(C:N) among capes and within-cape site differences for P. scouleri. Pairwise

comparisons were performed using Tukey’s HSD, alpha= 0.05.

Estimate Standard Error z-value Adjusted P-value Estimate Standard Error z-value Adjusted P-value

CP -0.016 0.005 3.134 0.020 FC vs. BB -0.022 0.007 -2.986 0.180

CA 0.018 0.009 2.068 0.292 MB -0.031 0.007 4.187 <0.01

CB -0.003 0.005 -0.584 0.992 BB vs. MB -0.009 0.008 1.152 0.999

CME 0.016 0.005 -2.906 0.040

PTA 0.067 0.005 -12.622 < 0.001

CA 0.034 0.009 3.892 0.001 YB vs. SH -0.013 0.007 -1.877 0.896

CB 0.013 0.005 2.812 0.052 TK -0.009 0.007 -1.323 0.996

CME 0.031 0.005 5.942 < 0.001 SH vs. TK 0.004 0.007 -0.509 1.000

PTA 0.083 0.005 -15.981 < 0.001

CB -0.021 0.009 2.456 0.130

CME -0.003 0.009 0.281 1.000

PTA 0.049 0.009 -5.525 < 0.001

CME 0.019 0.005 -3.651 0.003 CBN vs. CBS -0.008 0.007 1.123 1.000

PTA 0.070 0.005 -14.093 < 0.001 POH -0.044 0.008 5.806 <0.01

RP -0.063 0.007 9.055 <0.01

CBS vs. POH -0.035 0.008 4.637 <0.01

RP -0.054 0.007 7.750 <0.01

POH vs. RP -0.019 0.007 2.667 0.360

CME vs. PTA 0.051 0.005 -9.357 < 0.001 CMEN vs. CMES -0.008 0.020 0.393 1.000

KH 0.068 0.019 -3.514 0.038

MSP 0.026 0.020 -1.277 0.997

CMES vs. KH 0.075 0.008 -8.991 <0.01

MSP 0.034 0.010 -3.332 0.067

KH vs. MSP -0.042 0.009 4.880 <0.01

PTA MC vs. BML -0.055 0.006 -8.752 <0.01

CA vs.

CB vs.

Cape Site

CF vs.

CP vs.

Page 69: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

51

Table 2.7. Pairwise comparisons in log10(C:N) among capes and within-cape site differences for S. sessilis. Pairwise comparisons

were performed using Tukey’s HSD, alpha= 0.05.

Estimate Standard Error z-value Adjusted P-value Estimate Standard Error z-value Adjusted P-value

CP -0.006 0.003 1.750 0.475 FC vs. BB 0.005 0.006 0.961 1.000

CA 0.037 0.005 7.949 <0.001 MB 0.006 0.007 -0.855 1.000

CB 0.009 0.003 2.771 0.056 BB vs. MB 0.001 0.007 -0.138 1.000

CME 0.003 0.005 -0.507 0.995

PTA 0.013 0.007 -1.804 0.441

CA 0.042 0.005 9.240 <0.001 YB vs. SH -0.003 0.007 -0.390 1.000

CB 0.015 0.003 4.523 <0.001 TK -0.033 0.008 -4.337 <0.01

CME 0.008 0.005 1.676 0.525 SH vs. TK -0.031 0.006 5.106 <0.01

PTA 0.019 0.007 -2.606 0.087

CB -0.027 0.005 5.863 <0.001

CME -0.034 0.006 5.632 <0.001

PTA -0.023 0.008 2.893 0.040

CME -0.007 0.005 1.382 0.720 CBN vs. CBS -0.036 0.007 4.888 <0.01

PTA -0.004 0.007 -0.529 0.995 POH -0.011 0.007 1.454 0.985

RP -0.022 0.007 3.149 0.103

CBS vs. POH 0.025 0.009 -2.797 0.248

RP 0.014 0.008 -1.631 0.957

POH vs. RP -0.011 0.009 1.287 0.996

CME vs. PTA 0.011 0.008 -1.359 0.734 CMEN vs. CMES -0.009 0.012 0.710 1.000

KH 0.011 0.012 -0.946 1.000

MSP 0.002 0.014 -0.146 1.000

CMES vs. KH 0.020 0.010 -1.983 0.815

MSP 0.011 0.013 -0.827 1.000

KH vs. MSP -0.009 0.013 0.720 1.000

CA vs.

CB vs.

Cape Site

CF vs.

CP vs.

Page 70: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

52

Table 2.8. Effect of month and site on C:N in each species. Mixed effects models

included month and site as multiplicative fixed effects and species as a random intercept.

Species Source of variation df F-value P-value

F. distichus Month 1, 4772 5.999 0.014

Site 16, 4772 23.642 <.0001

Month: Site 16, 4772 6.001 <.0001

M. splendens Month 1, 1698 21.660 <.0001

Site 16, 1698 8.960 <.0001

Month: Site 16, 1698 3.180 <.0001

P. scouleri Month 1, 3924 5.670 0.017

Site 16, 3924 3.750 <.0001

Month: Site 16, 3924 3.370 <.0001

S. sessilis Month 1, 1558 2.980 0.085

Site 15, 1558 14.700 <.0001

Month:Site 15, 1558 4.030 <.0001

Page 71: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

53

Table 2.9. Effect of year and location (region, cape, and site analyzed separately) on C:N

for all species combined. Mixed effects models included year and location as

multiplicative fixed effects and species as a random intercept.

Source of variation df F-value P-value

Year 1, 4772 5.999 0.014

Site 16, 4772 23.642 <.0001

Year: Site 16, 4772 6.001 <.0001

Year 1, 4794 5.419 0.020

Cape 5, 4794 39.349 <.0001

Year: Cape 5, 4794 16.243 <.0001

Year 1, 4800 5.392 0.020

Region 2, 4800 61.237 <.0001

Year: Region 2, 4800 10.050 <.0001

Page 72: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

54

Table 2.10. Results from likelihood ratio tests to determine the most parsimonious model

fit for determining the influence of the environmental drivers N-availability, recent

upwelling, MEI, and NPGO on C:N. The full model included the multiplicative effects of

the four environmental drivers, which was compared to the model excluding the four-way

interaction between all environmental drivers to test the null hypothesis that the simpler

model is better. The outcome of the full model vs. 3-way interaction model for each

species was then compared with and without the first-order autoregressive correlation

structure (AR1), to account for autocorrelation present in the data.

Likelihood ratio P-value Likelihood ratio P-value

F. distichus 2.06 0.15 156.96 < 0.0001

M. splendens 0.23 0.63 285.5 < 0.0001

P. scouleri 0.005 0.94 19.22 < 0.0001

S. sessilis 22.8 < 0.0001 269.58 < 0.0001

Full model vs. 3-way interaction model: With correlation structure vs. without:

Page 73: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

55

Table 2.11. Results from mixed effects models on the influence of the environmental

drivers N-availability, recent upwelling, ENSO (multivariate ENSO index- MEI), and

NPGO on C:N. Model selection was performed on each species individually and for each

species there was a significant higher-order interaction among the four environmental

variables, shown here.

Species Interaction term df t-value P-value

F. distichus MEI: NPGO 673 -2.6 0.009

M. splendens N-availability: upwelling: MEI 1379 2.31 0.021

P. scouleri N-availability: MEI: NPGO 443 -1.99 0.046

S. sessilis N-availability: upwelling: MEI: NPGO 1290 -3.27 0.001

Page 74: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

56

Table 2.12. The independent contribution (%I) of each environmental driver in

explaining C:N for all species combined and each species individually. Percents are

based on the total R2 for each model.

Total R2 Species N-availability Recent upwelling MEI NPGO

All-species 0.531 97.213 0.358 2.022 0.123 0.284

F. distichus 0.055 -- 11.432 57.344 13.591 17.633

M. splendens 0.015 -- 78.586 4.953 10.347 6.115

P. scouleri 0.141 -- 53.475 19.986 3.818 22.722

S. sessilis 0.088 -- 24.724 65.986 0.898 8.392

Independent Contribution (%I)

Page 75: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

57

CHAPTER 3 – CONTRASTING DRIVERS OF GROWTH RATE IN TWO

DOMINANT ROCKY INTERTIDAL MACROPHYTES

ABSTRACT

Spatiotemporal variability in the growth rates of primary producers is a

fundamental aspect of ecological community structure, and understanding what drives

these patterns is an ongoing challenge for ecologists. Nutrient and light limitation are

often invoked as the primary drivers of these patterns. Through regulation of growth,

nutrient and light availability can influence tissue nutrient content in autotrophs, through

which community-level impacts can propagate. In rocky intertidal habitats, macrophytes

dominate the low intertidal zone and are major space occupiers, competitors, and primary

producers. Their patterns of growth therefore relate directly to their role in these

communities and the productivity of these critical habitats. Here, we investigate the

spatiotemporal patterns of the growth rates in two species that share similar roles as low

zone dominants: the kelp Saccharina sessilis, and the surfgrass, Phyllospadix scouleri,

across 900 km of coastline. Patterns in both the spatiotemporal variability in growth rates,

as well as the relationships to nutrient and light availability, differed for these two

species. Patterns of growth in S. sessilis were consistent with it being regulated by both

nutrient and light availability. Phyllospadix scouleri growth was also related to nutrient

and light availability, but the nature of the relationship depended on the availability of the

resource. Despite these findings, we did not observe a relationship between growth rates

and tissue nutrient content. By illustrating both the spatiotemporal patterns in growth and

Page 76: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

58

the relationship of growth to resource availability, these findings contribute to our

understanding of the bottom-up regulation of macrophyte assemblages in this system.

INTRODUCTION

A host of environmental, ecological, and physiological factors can affect plant

growth. Growth rates can be modulated by bottom-up influences, such as nutrients or

light, and by inter- or intra-specific interactions, such as competition for those resources

(Tilman 1985, Huston and DeAngelis 1994, Brauer et al. 2012). Understanding how

resources affect growth is critical to understanding the bottom-up regulation of

community structure. Further, growth can both influence, and be influenced by, how

macrophytes acquire and use nutrients (Ågren 2008), which can in turn play a role in

ecosystem functioning and community structure (Sterner and Elser 2002). By combining

growth measurements with our knowledge of spatiotemporal patterns in nutrient content

discussed in Chapter 2, we can start to assemble a more complete picture of the factors

that influence growth and how it interacts with nutrient content.

Limitation of growth rates in primary producers can occur via nutrient or light

limitation. The classic view of limitation by a single nutrient is Liebig’s Law of the

Minimum, wherein growth will be limited by the resource in lowest supply relative to the

plant’s needs (Liebig 1842). Research on limitation of plant growth has historically

focused primarily on nutrient limitation, leading to advances in our understanding of the

relationships between resources and productivity (LeBauer and Treseder 2008),

biodiversity (Harpole and Tilman 2007), and community structure (Carpenter et al.

2001). More recent theoretical work has shown, however, that when nutrient availability

Page 77: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

59

is high, light limitation is likely to regulate growth in single species models (Brauer et al.

2012). Limitation of plant growth, whether it is by nutrients or light, can have further

impacts through altering tissue nutrient content.

Primary producer tissue elemental content and the ratios between them have been

shown to play a role in ecosystem functioning and community structure across systems,

but there are fewer studies that look at their relationship to growth (Sterner and Elser

2002). In theory, elevated growth should increase demand for elements as the raw

material for the machinery of growth and metabolism. Of the primary limiting

macronutrients, phosphorus is required as the major component of ribosomal RNA

(rRNA), and nitrogen is required for protein synthesis (Sardans et al. 2012). As growth

declines, tissue carbon to nutrient ratios should increase, provided the decline in growth

is due to nutrient limitation (Sterner and Elser 2002). The identity of the limiting nutrient,

and its ambient concentrations, will influence this relationship, however. Much of the

foundational work in this field, particularly in aquatic habitats, has been conducted in

freshwater systems where the limiting nutrient is more commonly phosphorus (Sterner

and Elser 2002). In marine systems, nitrogen is generally considered to be the primary

limiting nutrient (Dugdale 1967), although more recent work has shown that in some

circumstances phosphorus can be limiting, or co-limitation can occur (Downing 1997,

Sundareshwar et al. 2003, Harpole et al. 2011). Further, in a nutrient-replete system

where nutrient availability may not impact growth as severely, instead allowing light

limitation to occur, the relationship between growth and nutrient content may change

Page 78: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

60

(Brauer et al. 2012). The relative availability of light and nutrients can influence growth

as well as tissue nutrient content (Sterner et al. 1997).

Applying this framework to our detailed knowledge of rocky intertidal

community and ecosystem dynamics should be informative. At the eastern boundaries of

most ocean basins, seasonal upwelling drives nutrient dynamics through the provision of

nutrient-rich deep waters to the nearshore. These eastern boundary upwelling systems are

highly productive on a global scale, especially when compared to the area they occupy

which is relatively small (Blanchette et al. 2009). On the west coast of North America,

upwelling occurs across the California Current Large Marine Ecosystem (CCLME), but

varies in strength and persistence latitudinally (Huyer 1983). Additional variation due to

coastal geomorphology and nearshore bathymetry arises at the meso- and local- scales.

Upwelling winds are strengthened at prominent coastal headlands (capes), leading to

stronger upwelling at these locations. Additionally, the width of the continental shelf

influences the nearshore retention of upwelled waters leading to the formation of large

phytoplankton blooms which attenuate light and may draw down nutrient concentrations

(Kavanaugh et al. 2009). These differences in upwelling and by extension in nitrogen

delivery to nearshore habitats (in the form of nitrate) are manifest in tissue nutrient

content, as we saw in Chapter 2, but the extent to which growth varies, what drives that

variation, and how it interacts with tissue nutrient content is unknown.

Growth of intertidal macrophytes can be influenced by factors ranging from

environmental influences, to ecological interactions, to physiological constraints.

Environmental drivers of growth in temperate intertidal seaweeds and seagrass can

Page 79: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

61

include nutrient availability (Aquilino et al. 2009), light availability (Dethier et al. 2005,

Kavanaugh et al. 2009), and temperature (Major and Davison 1998, Pfetzing et al. 2000).

Ecological interactions also frequently influence macrophytes in these systems.

Competition for resources, especially space, can be intense and has been found to

influence growth rates in a number of species (Worm and Chapman 1996, Dudgeon et al.

1999). Consumptive interactions also can influence growth rates. Herbivores are known

to selectively graze on tissues responsible for higher growth, photosynthesis, or nutrient

uptake (Bracken and Stachowicz 2007). On top of this selective grazing, tissue removal

in general can inhibit the ability of a macrophyte to grow simply by virtue of removing

biomass and surface area needed for photosynthesis and nutrient uptake.

At the individual level, physiological processes influence growth rates in

macrophytes. The process of photosynthesis is a major physiological mechanism through

which growth could be influenced. By providing the raw materials for photosynthesis and

plant growth to occur, nutrient and carbon uptake and assimilation are critical to the

process of growth as well. Nutrient uptake rates can depend on environmental factors

such as ambient nutrient concentrations (Naldi and Viaroli 2002), temperature (Lobban

and Harrison 1994), and tidal height (Bracken et al. 2011), as well as organismal or

physiological factors such as surface area to volume ratios, tissue type, and other aspects

of interplant variability (Lobban and Harrison 1994). Carbon uptake and assimilation can

also vary among species. Many intertidal macrophytes can utilize bicarbonate ions as a

carbon source via a carbon concentrating mechanism (CCM), but the extent to which they

do this may be driven by taxonomic differences (Maberly 1990), habitat or vertical

Page 80: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

62

distribution (Murru and Sandgren 2004), or other factors such as physiological limitations

or adaptations (Lobban and Harrison 1994). Interspecific differences in carbon and

nutrient uptake thus may relate to differences in growth patterns as well.

Combined, these complex drivers of growth in intertidal macrophytes have the

potential to inform our understanding of bottom-up regulation of these communities. The

low zone in temperate rocky intertidal habitats usually is typified by a dense macrophyte

community, and in the CCLME this includes two major space occupying species:

Saccharina sessilis and Phyllospadix scouleri. S. sessilis (C. Agardh) Kuntze is an

perennial kelp in the order Laminariales that is a dominant canopy-forming species in the

low intertidal (Dayton 1975). It grows from a basal meristem on a prominent holdfast and

can take on a bullate or strap-like form (Abbott and Hollenberg 1976). P. scouleri (W.J.

Hooker) is a perennial flowering seagrass that grows long blades from a dense rhizome

mat (Gibbs 1902). Phyllospadix spp. are known to have very high growth rates and are

some of the most productive communities worldwide (Ramirez-Garcia et al. 1998).

While these two species overlap some in their ecological roles (see above), they are very

different and are expected to follow different patterns of spatiotemporal variation and

relationships of growth to resource availability.

Here, we aim to characterize the growth patterns of these two dominant intertidal

species and investigate how nutrient and light availability influence growth patterns, and

how tissue carbon and nutrient content relates to macrophyte growth. Using observations

spanning multiple years and over 900 km of coastline in the CCLME, we relate known

Page 81: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

63

patterns in tissue nutrient content and environmental nutrient and light availability to

patterns in growth rate, addressing three primary research questions:

Q1: What are the patterns of spatiotemporal variability in growth rates in S.

sessilis and P. scouleri?

H1,1: Growth will vary seasonally, with highest growth in the months with the

most daylight hours (June).

H1,2: Growth will vary annually, with highest growth in years displaying

environmental conditions most favorable for growth (e.g. strong upwelling, La Niña).

H1,3: Growth will vary spatially, with highest growth in areas of higher nutrient

subsidies and light availability.

Q2: Do intertidal nutrient and light availability drive variation in growth rates?

H2,1: Growth rates will be greater at sites with higher nutrient and light

availability.

Q3: How do growth rates in S. sessilis and P. scouleri relate to carbon to nitrogen

ratios?

H3,1: As growth rates increase, carbon to nitrogen ratios should decrease signaling

elevated demand for nutrients. We expect this relationship to be stronger in P. scouleri, in

which we observed a possible indication of nutrient limitation (Chapter 2), compared to

S. sessilis, in which excess nutrient uptake is common.

H3,2: Tissue nitrogen content will be greater at sites with higher light availability,

due to higher demand from elevated growth.

Page 82: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

64

H3,3: Tissue carbon content will be greater at sites with higher light availability

due to higher rates of carbon fixation.

METHODS

Our study took place from April-September in 2008-2010 at 15 rocky intertidal

sites spanning approximately 900 km along the Oregon and northern California coasts

(Fig. 3.1). The spatiotemporal extent of sampling differed slightly by species, due to

availability of large populations of the target species at each site, with each species

sampled at 12 sites, 9 of which overlapped between the two species. Additionally, P.

scouleri growth rates used here only cover 2008 and 2009 and were only sampled May-

August.

Study Region

Our study sites are arranged in a hierarchical structure. Sites are replicates of

cape-level processes, and capes are grouped into regions that are broadly defined by

differing upwelling conditions (Menge et al. In Prep., Iles et al. 2012, Menge and Menge

2013). The northern Oregon region is typified by strong intermittent upwelling and

consists of Cape Foulweather (CF) and Cape Perpetua (CP). Within CF sites are (from

north to south): Fogarty Creek (FC), Boiler Bay (BB), and Manipulation Bay (MB; S.

sessilis only). Yachats Beach (YB), Strawberry Hill (SH), and Tokatee Klootchman (TK;

P. scouleri only) are the CP sites where we measured growth. The southern Oregon

region consists only of Cape Blanco (CB), a major upwelling center (Barth et al. 2000).

Sites within this cape are: Cape Blanco North (CBN), Port Orford Head (POH, P.

Page 83: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

65

scouleri only), and Rocky Point (RP). The northern California region includes Cape

Mendocino (CME) to the north and Point Arena (PTA), the southernmost cape. Sites

within CME include: Cape Mendocino North (CMEN; S. sessilis only), Cape Mendocino

South (CMES; S. sessilis only), Kibesillah Hill (KH), and MacKerricher State Park

(MSP). Finally, Point Arena sites included Moat Creek (MC; P. scouleri only) and

Bodega Marine Lab (BML).

Growth Measurements

We measured the growth rates of S. sessilis and P. scouleri in situ. Growth rates

of S. sessilis were sampled non-destructively using the hole punch method (Mann 1973).

At each site we used a 1 cm cork borer to punch a hole in each of 100 individual blades

of S. sessilis (not exceeding one blade per holdfast) at the center of the blade’s width and

approximately 5 cm above the holdfast. Since S. sessilis has a basal meristem, the hole

moves with the blade as it lengthens. The following low tide series (approximately two

weeks later), each area was resampled and all individuals that could be located were

measured (>50 individuals per sampling date). We measured the location from the base

of the blade to the bottom of each hole, and subtracted 5cm to account for the initial hole

punch placement in order to calculate growth. Rates were calculated on a mm per day

basis using the interval between sampling dates.

P. scouleri growth rates were measured using the destructive sampling methods

described in Zieman (1974). At each site and sampling date, a trio of P. scouleri sheaths

were bundled with a small cable tie and hole-punched with a small syringe needle just

below the top of the sheath in five replicate plots at a site. During the next low tide series,

Page 84: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

66

bundles of marked individuals were harvested. The growth of each new blade was

measured as the distance from the sheath to the pinprick hole.

Environmental drivers of growth rates

Intertidal nutrient availability (μmol/L NO3-1

) was sampled from each site on a

monthly basis. Samples were collected from a depth of approximately 1m and dissolved

inorganic nitrate was analyzed using the cadmium reduction method (Jones 1984). We

used monthly averages for each site and year in analyses. See Chapter 2 for additional

details.

Light availability was quantified as photosynthetically active radiation (PAR) in

the low zone of each of seven intertidal sites (cape abbreviations shown in parentheses):

Fogarty Creek (CF), Yachats Beach (CP), Cape Blanco North (CB), Port Orford Head

(CB), Kibesillah Hill (CME), Moat Creek (PTA), and Bodega Marine Lab (PTA). To

quantify light availability to intertidal macrophytes while immersed, when the majority of

production and growth occur, we measured downwelling PAR irradiance (400-700 nm)

to the intertidal zone to estimate the downwelling attenuation coefficient (Kd). Two

cosine corrected PAR sensors (LI-COR 190) were connected to a datalogger (Onset

Hobo) with a universal transconductance amplifier (EME Systems) and housed in a

custom, waterproof case with a window of spectrally neutral Plexiglas. The data logging

PAR sensors were leveled and attached to the rock, one in the low intertidal zone and

another above the intertidal zone. Light readings were recorded every 15 minutes. Data

from the two sensors were processed to eliminate out of water periods, nighttime readings

and any data records when the sun was at a zenith angle > 50° to avoid periods of high

Page 85: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

67

surface reflectance (Kirk 1994). The natural progression of the tides over the PAR sensor

in the low zone (PARlz) provided measurements of downwelling irradiance at different

water depths throughout the day. The proportion of incident light transmitted through the

water column was calculated using ln-transformed light data from both sensors as 1-

(PAR – PARlz)/PAR, and then regressed against the predicted depth of the water above

the tidal height of the PARlz sensor for all measurements recorded in a given day. The

average depth of the water was estimated to be the difference between predicted tidal

height and the elevation of the sensor (approximately 0- 0.1m above MLLW). The slope

of the regression is the daily estimate of the Kd of PAR, where higher values indicate

greater attenuation with depth of the sunlight incident on the water’s surface (and thus

less actinic light reaching benthic algae at a given water depth). For analyses, we

calculated the average Kd per day over the 15-day period prior to each growth

measurement.

Tissue carbon and nutrient content

To quantify elemental content of S. sessilis we collected 21 small (~1 in2) pieces

of tissue from the middle of the blades from 21 different individuals. For P. scouleri,

small handfuls of blades were taken from each of 3 areas where P. scouleri was growing

throughout a site. Samples were rinsed in the laboratory, sorted into three equal replicates

(each consisting of 7 tissue samples), dried, and ground into a fine powder. Percent

carbon and nitrogen were analyzed by the University of Georgia Analytical Lab in

Athens, Georgia, on a Carlo Erba NA1500 C/H/N Analyzer. See Chapter 2 for additional

details.

Page 86: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

68

Statistical Analysis

Analyses were conducted in R version 3.0.0 (R Core Team 2013). Spatiotemporal

patterns in growth rates were analyzed using mixed effects models with the package nlme

(Pinheiro et al. 2013) and all pairwise comparisons were performed using Tukey’s HSD

with alpha= 0.05 in the package multcomp (Hothorn et al. 2008). We included a first-

order autoregressive (AR1) correlation structure in these models to account for

autocorrelation within the data. All models were fit using restricted maximum likelihood.

We fit mixed effects model with the multiplicative fixed effects of month and year and

included the hierarchical spatial structure of sites nested within capes nested within

regions as a random intercept. We used Likelihood ratio tests to determine the most

parsimonious model fit. Pairwise comparisons were performed on the additive model for

both species. Because not all sites were sampled in all months and years we did not

include site in the above models to avoid spurious results. To look at spatial variation in

growth rates we fit mixed effects models as above but with region, cape, or site as the

fixed effect and the nested temporal structure of months within years as the random

intercept.

Because we are interested in the role of nutrient and light availability in

potentially limiting growth rates, we used quantile regression to assess the relationships

to growth across different quantiles of nutrients and light (Cade et al. 1999). Using the R

package quantreg (Koenker 2013), we fit models of the relationship between either

nutrient or light availability and growth rates (for each species separately) at twenty

different quantiles (τ; from 0.05 to 0.95 by increments of 0.05) of the explanatory

Page 87: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

69

variable (nutrients or light). To visualize these relationships we plotted the slope (β1) of

the linear fit (y = β0 + β1x + ε) across all quantiles (τ).

To analyze the relationship of growth rates to tissue carbon and nitrogen content,

we again used mixed effects models. Since we expect elemental content to influence

growth, here we used growth as a fixed effect, along with month, year, and cape as

multiplicative fixed effects, and site as a random effect. These models also included the

AR1 correlation structure. To select the most parsimonious model, we compared model

fits using Likelihood ratio tests. For P. scouleri, we only used data from 2009 to avoid

spurious results from uneven sampling of months and sites across years.

RESULTS

Spatiotemporal variability in growth rates

The pattern of monthly variation in S. sessilis growth differed depending on the

year (Fig. 3.2; month:year L-ratio= 34.13, P< 0.0001). Pairwise comparisons among

months revealed a seasonal pattern wherein growth was higher in April-July and lower in

August and September (Table 3.1; April = May = June = July > August = September).

Further, growth was similar in 2008 and 2009, but lower in both of those years compared

to 2010 (2009-2008: P-adj.= 0.37; 2008 and 2009 vs. 2010, P-adj.< 0.0001 for both).

We observed regional and cape-scale variation in growth rates of S. sessilis

(region: L-ratio= 21.99, P< 0.0001; cape: L-ratio= 26.17, P< 0.0001). Regional

differences separated by state, with the two Oregon regions not differing from each other,

but both having lower growth rates than northern California (Fig. 3.3; Table 3.2,

Northern OR vs. Southern OR: P-adj.= 0.90; P-adj< 0.0001 for both OR vs. CA

Page 88: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

70

comparisons). Capes fell out roughly along regional differences as well, with growth

rates at Cape Mendocino (CME) and Point Arena (PTA) being different from Cape

Foulweather (CF) and Cape Blanco (CB), but no other differences (Fig. 3.4, Table 3.3; P-

adj.< 0.01 for all).

Variation among sites in S. sessilis growth rates was also apparent (Fig. 3.5, Table

3.3; site: L-ratio= 67.29, P< 0.0001). From pairwise comparisons the two southernmost

sites were the primary drivers of the site effect: MacKerricher State Park (MSP, Cape

Mendocino), and Bodega Marine Lab (BML, Point Arena). MSP had the highest growth

rates and the most pairwise differences among sites, with higher growth than all sites

except Manipulation Bay (MB; P-adj= 0.92; vs. all other sites P-adj> 0.0001). BML, the

southernmost site, also had high growth rates, which were lower than those at MSP (P-

adj.= 0.01) but greater than those at FC and CBN (P-adj< 0.01 for both).

P. scouleri showed a different pattern of seasonal variation compared to S.

sessilis, although it should be noted that P. scouleri growth was not sampled in April as it

was in S. sessilis. Including the interaction between month and year showed no

improvement of model fit when compared to an additive model (Fig. 3,6, Table 3.4, L-

ratio= 2.43, P= 0.12). Growth rate in P. scouleri did show a seasonal pattern, with

differing growth rates among months, driven by lower growth in May compared to June

(F= 52.44, P< 0.0001; May-June P-adj.= 0.02). Growth rates of P. scouleri were lower in

2008 compared to 2009 (F= 222.52, P< 0.0001).

Spatial variation in P. scouleri was, at the regional-scale, similar to that in S.

sessilis. Both regions of Oregon had different growth rates compared to California;

Page 89: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

71

however, in P. scouleri growth was actually higher in Oregon than in California (Fig. 3.7,

Table 3.5, L-ratio= 20.81, P< 0.0001; Northern OR vs. Northern CA P-adj.= 0.007;

Southern OR vs. Northern CA P-adj.< 0.001; Northern OR vs. Southern OR P-adj.=

0.08). At the cape-scale growth rates at CME stood out from all other capes (Fig. 3.8,

Table 3.6; CME vs. CF P-adj.< 0.001, CME vs. CP P-adj.= 0.01, CME vs. CB P-adj.<

0.001, CME vs. PTA P-adj.= 0.02). Additionally, growth rates at CP and CB differed (P-

adj.= 0.02).

Variation among sites in P. scouleri was complex, with many site-level

differences (Fig. 3.9, Table 3.6; L-ratio= 81.1, P< 0.0001). There were, however, few

within-cape differences among sites, which we focus on here. At CP, YB had greater

growth rates than both SH and TK, which did not differ from each other (YB> SH= TK;

YB vs. SH and TK P-adj.< 0.01, SH vs. TK P-adj.= 1). And at PTA, MC had greater

growth rates than BML (P-adj.< 0.01).

Environmental drivers of growth rate

Nutrient availability varied among sites but, aside from Point Arena sites,

averages among capes were relatively similar (Fig. 3.10). Results from quantile

regression analyses on nutrient availability were varied. For S. sessilis, growth was

positively related to N-availability across most quantiles of N-availability except from the

70th

to the 90th, at which there was no relationship (Fig. 3.11a). At the 95th

quantile of N-

availability, S. sessilis growth was negatively related to N-availability. P. scouleri

growth, on the other hand, was positively related to N-availability at low levels of N-

Page 90: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

72

availability (5th

to 25th

quantiles), unrelated at the 30th

quantile, and negatively related

from the 35th

to 95th

quantile (Fig. 3.11b).

Differences in the influence of light attenuation on growth were also apparent

(Fig. 3.12). Cape Blanco and Point Arena, the two capes with replicate PAR sensors,

displayed contrasting light conditions between sites. Similarly to nutrient availability,

light availability was positively related to growth in S. sessilis and exhibited a more

complex relationship in P. scouleri (Fig. 3.13). In S. sessilis, the slope of the relationship

was consistently positive across quantiles of light attenuation from the 30th

to 55th

quantiles, and grew increasingly positive from the 65th

to the 85th

quantile (Fig. 3.13a).

Saccharina sessilis growth was unrelated to light attenuation at all other quantiles.

Phyllospadix scouleri growth was negatively related to light attenuation from the 40th

to

the 55th

quantiles, positively related from the 80th

to 95th

quantiles, and unrelated at all

other quantiles (Fig. 3.13b).

Relationships between growth and C:N

In S. sessilis, the relationship between elemental content and growth depended on

month, year, and cape for C:N, %C, and %N (Figs. 3.14 and 3.15, C:N L-ratio= 77.58,

P< 0.0001; %C L-ratio= 100.26, P< 0.0001; %N L-ratio= 54.99, P< 0.0001). In P.

scouleri, we did not observe any relationship between C:N, %C, and %N and P. scouleri

growth rates (Fig. 3.16). C:N, %C, and %N were influenced only by the interaction

between month and cape in our models (P< 0.004 for all).

DISCUSSION

Page 91: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

73

Our data indicate that two dominant intertidal macrophytes had contrasting

patterns and drivers of growth during upwelling season. Different spatiotemporal patterns

of growth in each species portend differences in environmental forcing, which were

evident in the relationship of each species to resource availability. Our evidence suggests

that the relationship of growth to resource availability (nutrients and light) differed

between species, with growth in S. sessilis increasing as nutrient availability increased

but decreasing as light availability increased. Growth in P. scouleri showed more

complicated relationships to nutrient and light availability, however, with relationships

varying across levels of resource supply. Despite these patterns, we saw little relationship

of growth to tissue nutrient content.

Contrasting spatiotemporal patterns in growth relate to environmental conditions

As expected, we observed seasonality in the growth rates of S. sessilis, which may

relate to seasonal patterns of light availability to the intertidal zone. Growth rates in S.

sessilis were approximately twice as high in March through July compared to rates in

August and September. As one of the dominant species in the low zone, rapid growth

may be necessary for this species to maintain competitive dominance (Dayton 1975), but

the late summer drop-off in growth rates was likely a result of environmental forcing. As

summer progresses on the Oregon and California coasts, the daily lowest low tides occur

progressively earlier in the morning. While evidence on the efficiency of photosynthesis

during immersion compared to emersion is mixed, most studies conclude that

photosynthetic gains are greater during emersion only if desiccation is minimal, a

challenge in intertidal habitats (Dring and Brown 1982, Madsen and Maberly 1990,

Page 92: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

74

Lobban and Harrison 1994, Williams and Dethier 2005). Given that, conventionally we

assume that photosynthesis is greater during immersion (Williams and Dethier 2005), and

that light availability measurements used here reflect underwater conditions. In the low

zone where S. sessilis occurs, decreased emersion times and splash from incoming waves

decrease the risks of desiccation. This may lead to higher photosynthesis during low tides

as well as immediately before and after when light attenuation by the overtopping water

column remains low. Thus, when low tides occur in the early morning before or at dawn,

as in August and September in this system, decreased light availability may explain the

decreased growth observed in S. sessilis in late summer. It is important to note, though,

that the scope of inference here is limited to the months we sampled, during the

upwelling season, and patterns may differ considerably at other times of year. The sample

period was chosen to capture dynamics during the growing season, as most macrophytes

stop growing and die back during the winter (personal observations).

Spatial patterns in S. sessilis growth are consistent with the observation that this

species’ growth is influenced by intertidal nutrient and light availability. Growth was

highest in California, a pattern that was largely consistent across spatial scales. At the

cape-level, growth in S. sessilis was lowest at Cape Perpetua. This cape is characterized

by a wide continental shelf across which upwelled waters are retained for long periods of

time, conditions that are favorable to the formation of large phytoplankton blooms (Chan

et al. 2008, Menge and Menge 2013). This result is consistent with Kavanaugh et al.

(2009) in which shading by phytoplankton was suggested to influence growth of S.

sessilis. Sites south of the Cape Mendocino sites (KH, MSP, and BML) all had high

Page 93: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

75

growth rates, driving the pattern of high growth rates of S. sessilis in California. These

sites are all characterized by high nutrient availability and relatively high light

availability, providing conditions that are favorable to growth.

While we did see a seasonal growth pattern in S. sessilis, as expected, we

observed few differences in growth rate among months in P. scouleri. The life history

and ecological role of P. scouleri may also play a role in the lack of seasonality of growth

patterns. P. scouleri is a perennial surfgrass, so steady growth throughout the summer

months may be more advantageous to its long-term persistence and competitive

dominance (Menge et al. 2005). Additionally, P. scouleri shows little seasonal variation

in cover (Turner 1985), so this steady growth may also be an aspect of the life history

strategy for this species.

In contrast to S. sessilis, P. scouleri grew faster in Oregon, also providing

evidence that growth is regulated differently in these two species, although the cape- and

site-scale patterns in this species were more complex. Despite having higher growth in

Oregon, at the cape level P. scouleri at Cape Perpetua had the lowest growth of all the

capes, as with S. sessilis. Nutrient and light availability were lowest of all capes at Cape

Perpetua. Spatial patterns of growth were consistent with the idea that nutrient and light

availability interact to drive P. scouleri growth. The highest growth rates, at the cape-

level, were observed at Capes Foulweather and Blanco, both in Oregon, where nutrient

and light availability were intermediate to other sites. At Point Arena, within-cape

variability among sites was pronounced. Growth of P. scouleri was very high at Moat

Creek and low at Bodega Marine Lab, both sites within Point Arena. These two sites

Page 94: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

76

contrasted in their nutrient and light conditions, with low nutrients and high light at Moat

Creek and high nutrients and low light at Bodega Marine Lab. Given that high nutrient

concentrations can cause toxicity in seagrasses (Larkum et al. 2006), these contrasting

conditions result in a more favorable environment for P. scouleri growth and account for

the large within-site variability in growth rates here.

For both species, interannual differences in growth rates were similar, with the

lowest growth rates in 2008, followed by 2009 and then 2010 (in the case of S. sessilis).

While we expected that growth rates would vary among years, we thought that these

differences would parallel known differences in oceanographic conditions that drive

nutrient subsidies to the rocky intertidal, namely upwelling and the El Niño Southern

Oscillation (ENSO). However, the trends we observed in growth across years are

basically opposite to what we would expect based on basin-scale oceanographic trends.

2008 and 2009 were characterized by strong upwelling, and in 2010 a brief El Niño event

occurred (McClatchie et al. 2009, Bjorkstedt et al. 2010, 2011). We expected growth to

be greater in times of strong upwelling, due to increased nutrient delivery to the

nearshore, and decrease during an El Niño event, due to decreased nutrient availability. In

fact, we saw the opposite. It is possible that high nutrient supply in years of strong

upwelling led to increased phytoplankton growth and decreased light availability in those

years, altering growth patterns in some locations. Further analyses perhaps on a longer

temporal scale may help elucidate these patterns.

Nutrient and light availability influence growth

Page 95: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

77

Relationships between growth and nutrient and light availability differed between

species. For S. sessilis, growth was positively related to nutrient availability across all but

the highest quantiles, that is, growth increased with increasing nutrient availability. In

contrast, S. sessilis growth decreased across most intermediate and high quantiles of light

attenuation, that is, as light availability increased (with decreasing light attenuation),

growth rates decreased. While this suggests that light played a role in regulating growth

in this species, the relationship between growth and light availability was opposite to our

expectations.

The relationship of P. scouleri to nutrient and light availability was more

complex. As nutrient availability increased, the rate of P. scouleri growth increased, but

at intermediate and high levels of nutrient availability, growth decreased. This is

consistent with nutrient toxicity, which has been documented in seagrasses (Larkum et al.

2006). In relation to light attenuation, there was a similarly varying pattern across levels.

At intermediate levels of light attenuation as light attenuation increased (light availability

decreased), growth decreased; the expected pattern. But, at the highest levels of light

attenuation (i.e. lowest light availability) we observed increasing growth rates. Combined

with insights from Chapter 2, these results indicate the potential for nutrient availability

to influence growth in this species, but a better understanding of where and when this

occurs will require further research. In particular, the potential for high nutrient

concentrations to inhibit P. scouleri growth or health is compelling and our evidence

suggests that may be a factor in this system.

Page 96: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

78

For both species, the relationships of growth rates to light availability differed

from our expectations. We observed increased growth rates at the lowest light levels (e.g.

high Kd), and for S. sessilis, there was an inverse relationship between light availability

and growth. We can interpret these findings by considering the various pathways through

which the intertidal light environment can be affected. During an upwelling event,

nutrient availability increases, fueling phytoplankton blooms, which attenuate light and

can shade macrophytes (Kavanaugh et al. 2009). This is particularly pronounced at

locations with a wide continental shelf, such as Cape Perpetua; however, we observed

increased, not decreased, growth at low light availabilities. From this chain of events, low

light conditions result when nutrient availability is high, fueling growth, and may be

reflected in our results. It is important to consider, also, that high light availability (i.e.

low Kd) can occur both when nutrient availability is high and nearshore retention is low

(e.g. upwelling), and when it is low (e.g. during downwelling). So, our observations that

the relationship of growth to light availability both differed from our expectations and

depended on the level of light availability (for P. scouleri especially) are likely related to

this complexity in oceanographic forcing of light availability.

Relation of growth to tissue nutrient content

Despite results indicating that nutrient and light availability influence macrophyte

growth, and theoretical and empirical evidence that this should effect tissue nutrient

content (Sterner and Elser 2002, Ågren 2008), we did not observe strong relationships

between tissue nutrient content and macrophyte growth in our study. For S. sessilis, the

relationship between C:N and growth depended on month, year, and cape. We know from

Page 97: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

79

Chapter 2 that patterns of spatiotemporal variation in C:N are complex. Here, we have

found that S. sessilis growth varies by month, year, and cape as well. So, the combination

of all of these into a four-way interaction reflecting C:N in this species is not surprising.

However, the relationships between C:N and growth rate in our study are weak. In P.

scouleri, C:N depended on the month, year, and cape; however, growth did not have any

bearing on C:N. Given that nutrient availability appeared to have a stronger relationship

to growth in this species, the lack of a relationship to C:N was not expected. The absence

of strong patterns relating tissue nutrient content to growth rates in these two species may

stem from multiple alternatives. Importantly, we did not measure growth and C:N on the

same individuals for this study, which likely impacted our ability to detect relationships

between these factors. Further, C:N data used in our analyses were matched to growth

data as monthly averages for each site in each year, so some temporal mismatch may be

masking patterns.

Broader implications for intertidal communities

This is one of the first studies to investigate spatiotemporal patterns of growth

rates, and the resources that drive them, across a biogeographic scale. For both species,

we found evidence that the dynamics of the upwelling process had some bearing on

macrophyte growth, whether through nutrient delivery or the more complex mechanism

of light attenuation. However, the evidence indicates that these dynamics varied between

species. The role of light availability in structuring intertidal communities is not often

considered, even though this resource varies in abundance over multiple spatial and

temporal scales. This study suggests that light availability has the potential to be a

Page 98: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

80

structuring mechanism in intertidal macrophyte assemblages. The insights gained here as

to the roles of nutrient and light availability on macrophyte growth clarify our

understanding of bottom-up regulation of community dynamics in rocky intertidal

habitats during upwelling season.

ACKNOWLEDGEMENTS

We wish to thank technicians at Oregon State University and Sonoma State

University, including (at OSU) Erin Richmond, Mae Noble, Gayle Murphy, Camryn

Pennington, Ryan Craig, Megan Poole, Lindsay Hunter, Angela Johnson, Jonathan

Robinson, Shawn Gerrity, Jeremy Henderson, and Mary Ellis all played integral roles in

collecting growth and C:N data. At SSU Adele Paquin, Russel DiFiori, Beatrice Cortina,

Terra Hazen, Rachel Francis, Samuel Briggs, Kandis Gilmore and Trevor Mangor all

spent considerable time and effort collecting and processing samples. We thank Adele

Paquin, Megan Wood and John Collins (SSU) for design and manufacture of custom

PAR cases and Adele Paquin and Megan Wood for maintenance of light sensors and light

data processing. We also thank the many interns and students who assisted with growth

measurements in the field and in the laboratory at both institutions. We also thank Susan

Williams (UC Davis, Bodega Marine Laboratory) for generously allowing the use of her

algal grinder. A. Barner, J. Rose, J. Sullivan, and C. Shen provided valuable feedback on

this manuscript. This research was supported by a National Science Foundation Graduate

Research Fellowship to S.L.C., NSF Award numbers OCE-0727611 and OCE-1061530

to K.J.N. and OCE-0726983 and OCE-1061233 to F.C., S.D.H., and B.A.M. and by the

Page 99: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

81

Partnership for Interdisciplinary Studies of Coastal Oceans (PISCO), funded by the David

and Lucile Packard Foundation and the Gordon and Betty Moore Foundation.

Page 100: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

82

FIGURES

Figure 3.1. Map of sampling locations, each point indicates one site, although some

points overlap. Growth rates of both species were measured at all sites labeled in bold.

All other sites either had only P. scouleri growth measured (italics) or S. sessilis (plain).

Sites are listed from north to south within a cape, and groupings by cape are as follows

(from north to south): Cape Foulweather (sites: Fogarty Creek (FC), Boiler Bay (BB),

Manipulation Bay (MB)); Cape Perpetua (sites: Yachats Beach (YB), Strawberry Hill

(SH), Tokatee Klootchman (TK)); Cape Blanco (sites: Cape Blanco North (CBN), Port

Orford Head (POH), Rocky Point (RP)); Cape Mendocino (sites: Cape Mendocino North

(CMEN), Cape Mendocino South (CMES), Kibesillah Hill (KH), MacKerricher State

Park (MSP)); and Point Arena (sites: Moat Creek (MC), Bodega Marine Lab (BML)).

Modified after Figure 2.1 in Chapter 2.

Page 101: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

83

Figure 3.2. Mean growth rate (mm/day) of S. sessilis across months, shown by year

(filled circles= 2008, open circles= 2009, filled triangles= 2010). Points show mean (±

SE).

Page 102: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

84

Figure 3.3. Growth rates vary among regions for S. sessilis. Northern Oregon consists of

Capes Foulweather and Perpetua, Southern Oregon includes only Cape Blanco, and

Northern California includes Cape Mendocino and Point Arena. Bars show means (± SE)

and different letters indicate differences among regions from Tukey’s HSD (alpha= 0.05).

Page 103: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

85

Figure 3.4. Variation in growth rates across capes for S. sessilis. Bars show means (± SE)

and different letters indicate differences among capes from Tukey’s HSD (alpha= 0.05).

Page 104: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

86

Figure 3.5. S. sessilis growth rate across sites, from north to south. Site codes are as follows (from north to south): Cape

Foulweather (sites: Fogarty Creek (FC), Boiler Bay (BB), Manipulation Bay (MB)); Cape Perpetua (sites: Yachats Beach (YB),

Strawberry Hill (SH)); Cape Blanco (sites: Cape Blanco North (CBN), Rocky Point (RP)); Cape Mendocino (sites: Cape

Mendocino North (CMEN), Cape Mendocino South (CMES), Kibesillah Hill (KH), MacKerricher State Park (MSP)); and Point

Arena (site: Bodega Marine Lab (BML)). Bars show means (± SE) and letters indicate significant pairwise comparisons among

sites from Tukey’s HSD (alpha= 0.05).

Page 105: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

87

Figure 3.6. Mean growth rate (mm/day) of P. scouleri across months, shown by year

(filled circles= 2008, open circles= 2009). Points show mean (± SE).

Page 106: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

88

Figure 3.7. Regional variation in growth rates for P. scouleri. Northern Oregon consists

of Capes Foulweather and Perpetua, Southern Oregon includes only Cape Blanco, and

Northern California includes Cape Mendocino and Point Arena. Bars show means (± SE)

and different letters indicate differences among regions from Tukey’s HSD (alpha= 0.05).

Page 107: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

89

Figure 3.8. Variation in growth rates among capes for P. scouleri. Bars show means (±

SE) and different letters indicate differences among capes from Tukey’s HSD (alpha=

0.05).

Page 108: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

90

Figure 3.9. P. scouleri growth across sites, listed from north (left) to south (right). Site codes are as follows (from north to south):

Cape Foulweather (sites: Fogarty Creek (FC), Boiler Bay (BB)); Cape Perpetua (sites: Yachats Beach (YB), Strawberry Hill

(SH), Tokatee Klootchman (TK)); Cape Blanco (sites: Cape Blanco North (CBN), Port Orford Head (POH), Rocky Point (RP));

Cape Mendocino (sites: Kibesillah Hill (KH), MacKerricher State Park (MSP)); and Point Arena (sites: Moat Creek (MC),

Bodega Marine Lab (BML)). Bars show mean (± SE) and letters indicate significant pairwise comparisons among sites from

Tukey’s HSD (alpha= 0.05).

Page 109: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

91

Figure 3.10. Nutrient availability ([NO3-1

] uM) averages for each site over the study period. Site codes are as follows (from north

to south): Cape Foulweather (sites: Fogarty Creek (FC), Boiler Bay (BB), Manipulation Bay (MB)); Cape Perpetua (sites:

Yachats Beach (YB), Strawberry Hill (SH), Tokatee Klootchman (TK)); Cape Blanco (sites: Cape Blanco North (CBN), Port

Orford Head (POH), Rocky Point (RP)); Cape Mendocino (sites: Cape Mendocino North (CMEN), Cape Mendocino South

(CMES), Kibesillah Hill (KH), MacKerricher State Park (MSP)); and Point Arena (sites: Moat Creek (MC), Bodega Marine Lab

(BML)). Bars show mean (± SE) and letters indicate significant pairwise comparisons among sites from Tukey’s HSD (alpha=

0.05).

Page 110: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

92

Figure 3.11. The slope of the relationship between N-availability and growth rate of (a) S.

sessilis and (b) P. scouleri across each quantile of N-availability from 0.05 to 0.95 by

increments of 0.05. Filled circles represent a significant relationship between the two

factors at each quantile at the alpha= 0.05 level; open circles represent a P-value greater

than 0.05.

Page 111: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

93

Figure 3.12. Light availability shown as the average downwelling attenuation coefficient

(Kd) for each site over the study period. Site codes are as follows (from north to south):

Cape Foulweather (site: Fogarty Creek (FC)); Cape Perpetua (site: Yachats Beach (YB)),

Cape Blanco (sites: Cape Blanco North (CBN), Port Orford Head (POH)); Cape

Mendocino (site: Kibesillah Hill (KH)); and Point Arena (sites: Moat Creek (MC),

Bodega Marine Lab (BML)). Bars show mean (± SE) and letters indicate significant

pairwise comparisons among sites from Tukey’s HSD (alpha= 0.05).

Page 112: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

94

Figure 3.13. The slope of the relationship between light attenuation coefficients (Kd) and

growth rate of (a) S. sessilis and (b) P. scouleri across each quantile of PAR from 0.05 to

0.95 by increments of 0.05. Filled circles represent a significant relationship between the

two factors at each quantile at the alpha= 0.05 level; open circles represent a P-value

greater than 0.05.

Page 113: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

95

Figure 3.14. S. sessilis %C, %N, and C:N across growth rates. %C, %N, and C:N values

are monthly averages for the same month, year, and site as growth measurements. Lines

show the fit of the model: elemental content= β0 + β1growth + ε. For all models, adjusted

R2 was less than or equal to 0.0006.

Page 114: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

96

Figure 3.15. The relationship between C:N and growth in S. sessilis across all three study

years and seasons (Spring= March-July, Summer= August-September). Data from each

cape are shown in different colors (blue= Cape Foulweather, red= Cape Perpetua, green=

Cape Blanco, purple= Cape Mendocino, orange= Point Arena).

Page 115: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

97

Figure 3.16. P. scouleri %C, %N, and C:N across growth rates. %C, %N, and C:N values

are monthly averages for the same month, year, and site as growth measurements. Lines

show the fit of the model: elemental content= β0 + β1growth + ε. For all models, adjusted

R2 was less than or equal to 0.0002.

Page 116: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

98

TABLES

Table 3.1. S. sessilis pairwise comparisons among months. Comparisons were performed

using Tukey's HSD, alpha= 0.05.

Estimate SE z-value Adjusted P-value

May 0.972 0.811 -1.199 0.810

June 0.588 0.805 -0.731 0.973

July 0.698 0.804 -0.868 0.944

August 2.059 0.805 -2.559 0.088

September 2.677 0.821 -3.260 0.011

June -0.384 0.174 2.211 0.198

July -0.274 0.171 1.602 0.552

August 1.087 0.179 -6.067 <0.001

September 1.705 0.255 -6.690 <0.001

July 0.1096 0.1244 -0.881 0.9408

August 1.4701 0.132 -11.141 <0.001

September 2.0884 0.2328 -8.971 <0.001

August 1.3605 0.1228 -11.079 <0.001

September 1.9788 0.2282 -8.67 <0.001

August vs. September 0.6183 0.2316 -2.669 0.0658

April vs.

May vs.

June vs.

July vs.

Month

Page 117: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

99

Table 3.2. Regional differences in growth rates of S. sessilis. Comparisons were

performed using Tukey's HSD, alpha= 0.05.

Estimate SE z value Adjusted P-value

Northern OR vs. Southern OR 0.056 0.125 -0.448 0.895

Northern CA -0.611 0.143 -4.270 <0.0001

Southern OR vs. Northern CA -0.667 0.155 -4.291 <0.0001

Region

Page 118: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

100

Table 3.3. Cape- and site-scale differences in growth rates of S. sessilis. Pairwise comparisons were performed using Tukey’s

HSD, alpha= 0.05.

Estimate SE z-value Adjusted P-value Estimate SE z-value Adjusted P-value

CP -0.322 0.159 2.026 0.246 FC vs. BB -0.520 0.240 -2.166 0.530

CB -0.126 0.154 -0.817 0.923

CME -0.762 0.187 4.081 < 0.001

PTA -0.805 0.225 3.574 0.003

CB 0.196 0.143 1.373 0.638 YB vs. SH 0.093 0.198 0.468 1.000

CME -0.440 0.185 -2.384 0.115

PTA -0.483 0.222 2.170 0.185

CME -0.636 0.179 3.554 0.003 CBN vs. RP -0.370 0.193 1.917 0.712

PTA -0.679 0.217 3.123 0.015

CME vs. PTA -0.042 0.241 0.176 1.000 CMEN vs. CMES -0.345 0.536 0.644 1.000

KH -0.416 0.448 0.930 0.998

MSP -2.020 0.463 4.362 <0.01

CMES vs. KH -0.072 0.451 0.159 1.000

MSP -1.675 0.466 3.590 0.014

KH vs. MSP -1.603 0.325 4.927 <0.01

Cape

CF vs.

CP vs.

CB vs.

Site

Page 119: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

101

Table 3.4. P. scouleri pairwise comparisons among months. Comparisons were

performed using Tukey's HSD, alpha= 0.05.

Estimate SE z-value Adjusted P-value

June -0.817 0.287 2.847 0.022

July -0.749 0.300 2.495 0.059

August -0.630 0.304 2.072 0.158

July 0.068 0.218 -0.313 0.989

August 0.186 0.227 -0.821 0.841

July vs. August 0.118 0.207 -0.570 0.940

May vs.

June vs.

Page 120: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

102

Table 3.5. Regional differences in growth rates of P. scouleri. Comparisons were

performed using Tukey's HSD, alpha= 0.05.

Estimate SE z value Adjusted P-value

Northern OR vs. Southern OR 0.462 0.216 2.134 0.083

Northern CA 0.621 0.208 2.988 0.008

Southern OR vs. Northern CA 1.083 0.238 4.546 < 0.001

Region

Page 121: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

103

Table 3.6. Cape- and site-scale differences in growth rates of P. scouleri. Pairwise comparisons were performed using Tukey’s

HSD, alpha= 0.05.

Estimate SE z-value Adjusted P-value Estimate SE z-value Adjusted P-value

CP 0.701 0.262 -2.676 0.057 FC vs. BB -0.329 0.430 -0.765 1.000

CB -0.009 0.273 -0.032 1.000

CME 1.672 0.327 -5.112 <0.001

PTA 0.647 0.300 -2.157 0.193

CB -0.710 0.233 -3.043 0.020 SH 1.314 0.367 3.581 0.017

CME 0.971 0.301 3.230 0.011 TK 1.337 0.358 3.739 <0.01

PTA -0.054 0.263 0.205 1.000 SH vs. TK 0.023 0.362 -0.062 1.000

CME 1.681 0.307 -5.485 <0.001 CBN vs. POH 0.444 0.434 -1.024 0.997

PTA 0.656 0.272 -2.411 0.110 RP 0.541 0.388 -1.397 0.963

POH vs. RP 0.097 0.454 -0.214 1.000

CME vs. PTA -1.025 0.332 3.090 0.017 KH vs. MSP 0.205 0.504 -0.408 1.000

MC vs. BML 2.208 0.448 4.924 <0.01

Cape

CF vs.

CP vs.

CB vs.

Site

YB vs.

Page 122: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

104

CHAPTER 4 – AN EXCEPTION TO THE RULE? ABIOTIC FACTORS SET THE

LOWER LIMIT OF THE HIGH INTERTIDAL SEAWEED FUCUS DISTICHUS

ABSTRACT

Patterns of zonation in ecological systems arise due to different range limits

among species, which are typically determined by interacting abiotic and biotic factors.

How these patterns are set is a long-standing question in ecology, and understanding

exceptions to current models of these processes is critical under changing climate

conditions. In rocky intertidal habitats, vertical zonation occurs over a narrow elevation

range between low and high tide. This elevation range spans extreme environmental

gradients in desiccation stress, heat stress, and light availability associated with

immersion and emersion time. The classic conceptual model of rocky intertidal zonation

is that biological factors (e.g. herbivory, predation, competition) set the lower limits of a

species’ vertical distribution and abiotic factors (e.g. desiccation, heat stress) set the

upper limits. We tested this model using the common rocky intertidal seaweed, Fucus

distichus, a high zone species. The null hypothesis was that biological factors set its

lower distributional limit, and our alternative hypothesis was that the limit was set by

physical or physiological factors. We performed reciprocal transplants of F. distichus

individuals in the field, evaluating four factors that were hypothesized to set the lower

limits for this species: immersion stress, canopy competition for space, light, and/or

nutrients, and herbivory. Field manipulations revealed that F. distichus did poorly in the

low zone, regardless of the presence of a macrophyte canopy or herbivores, suggesting

abiotic factors play a role. Laboratory mesocosm experiments tested two proposed

Page 123: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

105

mechanisms for the intolerance of F. distichus to low zone conditions: immersion time

was too long and/or light availability was too low. Results suggested that reduced light

availability was the underlying mechanism for poor performance of F. distichus in the

low zone. To our knowledge, this is the first experimental evidence that light and not

submergence determines the lower limit of F. distichus. Taken together, these results

suggest a role of physical factors in setting the lower distributional limit in this species,

providing an exception to the common model of rocky intertidal zonation.

INTRODUCTION

Species range limits are restricted by environmental conditions and species

interactions, but determining the mechanisms underlying range limits remains a

fundamental challenge in ecology. Classically, we consider a species’ range to be the

balance between physically and biologically determined limits, with environmental stress

setting one limit to this range and, often in the more environmentally benign part of the

range, species interactions setting the other limit (e.g. Connell 1972). Evidence indicates

that these general patterns hold across many ecological systems. For example, in alpine

habitats, the upper distribution of vegetation along an elevation gradient is generally set

by thermal tolerances, whereas more complex ecological dynamics, including

competition, play a larger role in mediating lower elevation community structure (Reiners

and Lang 1979, Tang and Ohsawa 1997). Benthic invertebrate densities as well as fish

populations tend to follow zonation patterns in riverine and stream habitats depending on

flow and water temperature (Statzner and Higler 1986). In a Rocky Mountain stream

system, Rahel and Hubert (1991) found that zonation of fish was determined by a

Page 124: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

106

combination of physical and biotic factors. At upstream reaches, fish assemblages were

dependent on thermal tolerances with cold-water species thriving, whereas downstream

reaches were mainly influenced by the addition of new species and associated species

interactions. And in marine rocky intertidal systems the classic paradigm is that upper

intertidal range limits are set by environmental stress and lower limits by predation or

competition (Connell 1961, Paine 1966, Lubchenco 1980).

At different levels of environmental stress different, types of interactions are

expected to influence species and drive community regulation, as hypothesized by

environmental stress models (ESMs) (Menge and Sutherland 1987, Bruno et al. 2003). At

low to intermediate levels of environmental stress, species interactions are proposed to be

the mechanisms that determine community structure. At high and very high levels of

environmental stress, positive species interactions such as facilitation, as well as stress

itself, are hypothesized to be the important processes influencing species and the

community. Thus, across gradients of environmental stress a suite of abiotic and biotic

factors interact to determine species ranges and community structure.

Rocky intertidal habitats have steep gradients in environmental conditions across

relatively small (vertical) spatial scales, e.g. on the order of tens of meters, making it a

tractable system in which to experimentally investigate the determinants of species’ range

limits. Environmental gradients are particularly pronounced going vertically (i.e.

perpendicular to the water’s edge) from the lower intertidal (low zone) to the upper

intertidal (high zone) (Lewis 1964). At low tide, areas higher on the shore are subject to

greater desiccation and heat stress than areas closer to the water, due to longer emersion

Page 125: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

107

times. At high tide, organisms in the low zone are submerged longer, and thus have

longer to search for prey (mobile consumers), or to feed/obtain nutrients and grow (filter

feeders and autotrophs) and thus increase their competitive ability. In macrophytes,

longer times underwater may also result in less light exposure, potentially limiting these

light-dependent species.

Rocky intertidal zonation is widespread and, although the actual species

composition varies, the overall pattern of zonation is remarkably consistent on a global

scale, at least in temperate regions (Stephenson and Stephenson 1949). Typically, the low

zone has high cover of macrophytes and high abundance of mobile invertebrates, the mid

zone is dominated by bivalves (mussels or oysters), and the high zone typically has lower

overall cover of organisms, which consist mostly of smaller macrophytes, barnacles, and

other smaller invertebrates. Foundational work in this system determined that the upper

limits of intertidal species distributions typically are set by abiotic factors, while lower

limits are set by biotic interactions (Connell 1961, Paine 1966, Lubchenco 1980). These

and other studies led to the paradigm that conditions in the high zone are more stressful

for intertidal organisms, and that the low zone is more benign and thus likely to be the

preferred habitat for most species. Pairing this understanding of zonation with the ESMs,

community regulation in the low zone then is more likely to be driven by predation or

competition (negative interactions). The high zone is assumed to be a more stressful

environment, with stress, as well as positive interactions such as facilitation, expected to

serve as the dominant structuring forces.

Page 126: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

108

Implicit within this model of rocky intertidal zonation is the assumption that the

same conditions will be stressful to all intertidal species or individuals, and specifically

that the high zone is the more stressful environment in this habitat. Yet, our conceptual

models of the drivers of intertidal zonation may not adequately account for among-

species variation in adaptations to life under high zone conditions. High zone species can

be adapted to the physical conditions of the high zone, so they may be tolerant of, or even

require, these conditions (Hays 2007). Further, longer periods of submersion and the

lower light availability of the low zone may actually pose environmental stress to an

organism adapted to high zone conditions.

The literature testing hypotheses of intertidal species’ limits goes back more than

a century (Baker 1909), but not all of it has confirmed the paradigm that lower limits are

set by biotic interactions. For intertidal seaweeds especially, the drivers of zonation

appear to be more complex. Some studies have found coherence with the classic model of

zonation drivers (Lubchenco 1980, Harley 2003), others demonstrate inconsistencies

(Druehl and Green 1982, Stengel and Dring 1997, Dethier and Williams 2009). For

instance, consistent with the model, Lubchenco (1980) found, in an experimental

investigation of zonation of intertidal macroalgae in New England, USA, that the lower

boundaries of the dominant mid-intertidal species Fucus distichus and F. vesiculosus

were set by competition with the low zone dominant macroalga Chondrus crispus, while

desiccation tolerance set their upper limits. In contrast, other studies have uncovered

more complicated dynamics. In their study of F. distichus, Dethier and Williams (2009)

found that the seaweeds performed best in the mid zone in the spring and the high zone in

Page 127: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

109

the fall, thus the zone of most stress shifted seasonally. They attributed these results to

seasonal variation in both cumulative immersion time and time of low tides (shifting

from daytime to nighttime), suggesting the potential role of light limitation, in addition to

immersion stress, in causing a stress response at low tidal heights. In this case, and

others, patterns of immersion and emersion have been invoked as additional or alternative

drivers of zonation, especially for high zone autotrophs (Baker 1909, Druehl and Green

1982, Dethier and Williams 2009).

Understanding how environmental stress acts to determine species niche widths is

a basic question in ecology, and one that relates more urgently to the issue of predicting

species distributions under future climate change scenarios. The documented examples

of, and exceptions to, the paradigm of rocky intertidal zonation play an important role as

the basis for forecasting the impact of climate change on intertidal communities (Helmuth

et al. 2005, Harley et al. 2006, 2012). Qualitative predictions are largely based on the

understanding that lower limits are set by biotic interactions and upper limits are set by

abiotic factors. Such predictions are usually qualified by noting that potential responses

are highly uncertain and complicated (Harley et al. 2006, 2012). In general, the vertical

ranges of many intertidal species are expected to be “squeezed” by climate change effects

as sea level rise pushes lower limits higher on the shore, and increasing air temperatures

push upper limits lower.

Here we investigate drivers of the vertical distribution of a high-zone rocky

intertidal seaweed, Fucus distichus subsp. evanescens (C. Agardh) Powell (hereafter:

Fucus), asking: What factor(s) set(s) the lower limit for this high zone species? We first

Page 128: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

110

performed a field reciprocal transplant experiment, testing three hypotheses of factors

that may operate either in isolation or in concert to limit Fucus’s vertical distribution:

H1) Immersion stress (“drowning”) sets the lower limit for Fucus;

H2) Fucus is out-competed in the low zone by other canopy-forming macrophytes,

thus canopy competition sets its lower limit;

H3) Herbivory is greater in the low zone, thus Fucus is consumed in the low zone.

We also took field observations to assess two additional hypotheses on factors

that might limit Fucus’s vertical distribution from a different life stage, as recruits:

H4) Fucus recruits are dispersal-limited and thus do not reach the low zone; and

H5) Fucus requires high primary cover of barnacles, which is typically found

higher on the shore because barnacles facilitate Fucus recruitment.

We predicted that the factors regulating Fucus’s vertical distribution would be

consistent with the classic model of rocky intertidal zonation in that biotic interactions

(competition, herbivory) would be more important than immersion stress in setting the

lower limit of Fucus’s distribution. Based on research that has shown that intertidal

macrophytes tend to have short dispersal distances (van Tamelen et al. 1997, Shanks et

al. 2003), we also predicted that that recruitment to the low zone is likely limited, and

thus recruit density would be greatest higher on the shore.

To test our hypotheses, we used a combination of field observations and

experiments, and laboratory mesocosm experiments. A field experiment was designed to

directly test hypotheses H1-H3; field observations were taken to evaluate hypotheses H4

and H5, and the mesocosm experiment was done to further test H1, specifically the

Page 129: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

111

mechanisms underlying differences in Fucus performance in the high and low zone in

field experiments. That is, the lab experiment tested two factors that we hypothesized

accounted for the variation in Fucus performance in high vs. low zone transplants:

immersion time and light level.

METHODS

Study species

Fucus distichus subsp. evanescens (C. Agardh) Powell (hereafter: Fucus) is a

common brown alga found throughout the west coast of North America from Alaska to

Central California and in other areas of the Pacific and Atlantic oceans (Gabrielson et al.

2006, Guiry and Guiry 2010, Lindeberg et al. 2010). Fucus plays an important role in

ecological interactions in the high zone in Oregon (Farrell 1991), and members of this

genus are found throughout northern temperate rocky intertidal regions (Wahl et al.

2011). On the Oregon coast, Fucus is found in the mid-high to high zone. Individuals

grow from a small discoid holdfast in a dichotomous branching pattern, so growth can be

measured in situ by total length and the number of tips greater than 1cm from the last

dichotomy (Dethier et al. 2005). Additionally, reproductive structures (receptacles) form

at the tips and can be enumerated to measure reproductive status (Abbott and Hollenberg

1976, Dethier et al. 2005).

Study Site

The field portion of our study was conducted at Fogarty Creek (4450’24” N,

124 3’36” W) on Cape Foulweather, a major rocky headland along the central Oregon

Page 130: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

112

coast. The experiment ran for three months, April-July, 2011. Fogarty Creek (FC)

experiences strong intermittent upwelling and because of the narrow continental shelf at

this cape the upwelled waters are not retained close to shore for long periods. Thus, the

waters off of FC are characterized by high nutrient availability and relatively low light

attenuation in part due to relatively low phytoplankton biomass (chlorophyll a) (Menge et

al. In Prep., Kavanaugh et al. 2009, Menge and Menge 2013). At FC, Fucus is common

in the mid-high to high zone, and is rarely found either within or below the mussel bed.

While Fucus has been shown to have varying morphologies at this site based on wave

exposure (Blanchette 1997), our entire study took place within the exposed area of the

site, limiting our inference to those conditions.

Field tests of vertical limitation

To test the effects of tidal height on Fucus, we performed a reciprocal transplant

field experiment. Following the methods of Blanchette (1997) we transplanted

individuals to the low zone (+0.15m above MLLW; below the mussel bed) and the high

zone (+1.1m above MLLW; Fucus’ typical tidal height). Using hammers and chisels, we

first removed 128 Fucus individuals, including the rock around each holdfast. To reattach

them in their new locations, we chiseled a small depression in the substrate and used

marine epoxy putty (Z-spar A-788 Splash Zone Epoxy, Pettit Paint Company) to affix the

rock to the substrate. Transplants were arranged in sets of two, to insure against loss of

experimental individuals during the course of the experiment as can frequently occur in

Fucus and other fucoids (Blanchette 1997, Hays 2007). In each zone we set up four

Page 131: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

113

treatments with eight replicates each: canopy removal, herbivore removal, paint control,

and an unmanipulated control.

We manipulated herbivore densities around Fucus transplants in both zones to test

the hypothesis that herbivory limits the vertical distribution of Fucus. To exclude

herbivores, we laid a barrier of marine epoxy putty painted with copper anti-fouling paint

in a 25x25cm (.0625 m2) square around the Fucus transplants. Molluscan grazing

invertebrates (limpets, chitons) are deterred by a copper paint barrier (Cubit 1984,

Benedetti-Cecchi and Cinelli 1997, Menge et al. 1999, Freidenburg et al. 2007, Guerry

2008). Realistically, however, recruitment by limpets and the occasional entry of a few

adult individuals make the exclusion treatment a reduction in herbivory treatment. Once

the copper paint barrier was set up, we counted and removed all herbivores from within

the barrier in the herbivore removal treatments, and maintained the treatment manually in

this way every two weeks. Herbivores in the paint control treatments were counted every

two weeks, but not removed. At the end of the experiment, mean herbivore densities in

high zone paint controls were 41.6 (± 8.0) vs. 11.1 (± 4.3) in the herbivore removals

(mean ± SE). In the low zone, mean herbivore densities in paint controls were 4.9 (± 1.0)

vs. 4.5 (± 1.3) in the herbivore removals (mean ± SE).

We used three controls in this experiment: marked naturally-growing individuals,

unmanipulated controls, and paint controls. To control for effects of transplantation itself

we followed the growth of 30 marked but undisturbed Fucus individuals in the high zone

(marked individuals). In unmanipulated controls Fucus individuals were transplanted to

the high and low zones, but had no other manipulations. In paint controls, we controlled

Page 132: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

114

for the effects of the copper paint barrier used in the herbivore removal treatment by

applying marine epoxy and copper paint only to the corners of a 0.0625m2 square around

the transplants.

To assess the hypothesis that competition with other canopy-forming macrophytes

limits Fucus’ vertical distribution, we manipulated macrophyte canopy adjacent to

transplants. In canopy removal treatments we manually removed any adjacent seaweeds

that were large enough to come into contact with or shade the Fucus transplants and

maintained these removals over the course of the experiment. The unmanipulated

controls served as the control treatment for this experiment.

In our field experiment, Fucus individuals were measured every low tide series

(15 ± 1.2 days between measurements, mean ± SE), and they were measured weekly in

the laboratory mesocosm experiment. For each individual we measured the total length

(cm), number of tips >1cm, and number of tips that had formed receptacles. In the lab

experiment we also measured wet weight (g) after blotting with a towel. Here we report

results as a change in weight (mesocosms only), length, number of tips, or the number of

receptacles from the beginning to the end of each experiment (87 days for the field

experiment and 32 days for the mesocosm experiment). We refer to these metrics from

here on as growth in weight, growth in length, new tip growth, and new receptacle

formation, respectively. We also report the final receptacle proportion, which is the

quotient of the final number of receptacles divided by the final number of tips.

We used observational methods to assess the hypotheses that (H4) the vertical

distribution of Fucus is limited by the distance Fucus recruits can travel, and (H5) the

Page 133: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

115

presence of barnacles will facilitate recruit attachment to the rock. For this component,

we sampled along three vertical transects from the area of low zone transplants to the

area of high zone transplants three times during the course of the experiment. We placed

a 0.25m2 quadrat at every other meter along each transect within which we estimated the

percent cover of Fucus adults, primary cover of barnacles (Balanus glandula and

Chthamalus dalli [combined in observations], and Semibalanus cariosus), and secondary

cover of barnacles (primarily attached to the mussel Mytilus californianus). We also

counted all fucoid recruits within each quadrat. Because Fucus and the other high

intertidal fucoid alga that is common in this region, Pelvetiopsis limitata, can be difficult

to distinguish at very small sizes, we cannot be certain that all counted recruits were

Fucus.

Mesocosm tests of vertical limitation

To disentangle the role of two of the major physical factors that differ between the

high and low zones, and their effect on Fucus growth, we conducted a two-way factorial

mesocosm experiment manipulating immersion time and light level at Hatfield Marine

Science Center, Newport, Oregon. Two immersion-time treatments were chosen to mimic

the immersion/emersion patterns of the high and low zone treatments in the field portion

of this study (+0.15 and +1.1 m above MLLW, respectively). To test the effect of light,

half of the tidal mesocosms were covered with shade cloth (shade) and half were left

uncovered (full). All mesocosms were randomly assigned to a treatment combination.

Each treatment combination was replicated five times, totaling twenty mesocosms, and

each mesocosm held 19 liters of seawater.

Page 134: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

116

The mesocosms were constructed in an indoor cold room lab space by plumbing

the mesocosms to alternately fill and drain using solenoid valves attached to 7-day digital

timers. These timers were programmed to open the solenoid valves, thus draining the

mesocosms, using tide tables from the National Oceanic and Atmospheric Administration

(NOAA) Tides and Currents website (tidesandcurrents.noaa.gov) to determine aerial

exposures expected at +0.15m and +1.1m above MLLW, respectively, after which time

the solenoids closed to allow the next submersion period. Full spectrum fluorescent lights

set to be on from 8am to 5pm daily were installed above the mesocosms. Average (± SE)

light levels out of water in the full treatments were 46.75 (± 2.61) µmol photons m-2

s-1

and 15.25 (± 1.16) µmol photons m-2

s-1

in the shade treatments. While submerged,

average light levels were 15.74 (± 2.66) µmol photons m-2

s-1

in the full treatments and

5.47 (± 0.39) µmol photons m-2

s-1

in the shade treatments. Light was measured using a

light meter and underwater quantum sensor (LiCor Biosciences, Lincoln, Nebraska, L-

192 SA).

To initiate the mesocosm experiment, we chiseled 60 Fucus individuals, including

some rock attached, from the same area of FC as in the field portion of this study. We

made sure to leave some rock attached to the holdfasts so that they would stand upright in

their natural position, and not float when submerged. Three individuals were marked with

different colored small cable ties and placed in each mesocosm at the start of the one-

month experiment. At the beginning and end of the experiment, and each week in

between, we measured the wet weight (after blotting with a towel), length, number of tips

>1cm, and number of receptacles. In analyses we corrected weights by subtracting the

Page 135: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

117

weight of the rock and cable tie (taken at the end of the experiment) from each weight

measurement so that any changes reflected changes in Fucus growth or tissue loss.

Physiological measurements

Photosynthesis versus irradiance measurements were made over the course of two

days at end of the mesocosm experiment. We measured photosynthesis on 20 Fucus

individuals: one randomly selected from each of the 20 mesocosm replicates. After being

given 10 minutes to acclimate to the initial irradiance, each Fucus individual was

incubated in a closed 7.5-liter Plexiglas cylindrical chamber for 10 minutes at each of

four different levels of irradiance: 153, 633, 1137, 1769 µmol photons m-2

s-1

. Water was

circulated in each chamber with a circulation pump (Aqueon ACP 500, Aqueon Products,

Franklin, WI) in order to avoid the formation of boundary layers. At each irradiance

level, the system was closed and dissolved oxygen and temperature were measured at the

beginning and end of a ten minute interval. Between each irradiance measurement, the

system was flushed with water. After all light levels were tested, the chambers were

covered and dark respiration rates (at 0 µmol photons m-2

s-1

) were measured over the

course of sixty minutes.

We calculated three of the photosynthetic parameters from a linear fit between the

rates measured at light levels 0 (respiration) and 153 µmol photons m-2

s-1

for each

replicate run of photosynthesis versus irradiance: respiration rate (respiration),

compensation irradiance (Ec, irradiance at which photosynthesis equals zero), and slope.

The fourth parameter was the maximum photosynthetic rate for each trial. From these

Page 136: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

118

parameters we fit Michaelis-Menten relationships to the photosynthesis versus irradiance

plots to visually compare treatments (Tait and Schiel 2011).

Carbon and nitrogen content were analyzed on a subset of individuals from both

field and lab experiments at the time of completion (nfield= 24, nlab= 20). Samples were

rinsed with nanopure water, dried at 60°C for 48 hours, ground to a fine powder on a

Spex Sigma Prep 8000D Mixer/Mill, and stored in ashed glass vials. Percent C and N

were analyzed by the Boston University Stable Isotope Laboratory on a Fisons NA 1500

elemental analyzer and measured using Eager200 software. Check standards are inserted

into the run to ensure precision and quality control, and any anomalous samples are

reweighed and rerun.

Statistical Analyses

All data analyses were carried out using R, version 3.0.0, and packages

AICcmodavg (to calculate AICc), and car (Fox and Weisberg 2011, Mazerolle 2013, R

Core Team 2013).

Due to an unbalanced dataset resulting from loss of experimental individuals over

the course of the field experiment, we fit generalized linear models (GLM) to field

experimental data. All continuous responses were fit to a GLM with a Gaussian

distribution using the identity link function. Final receptacle proportion was fit to a GLM

with a binomial distribution and the logit link function. To first assess the immersion

stress hypothesis, we performed drop-in-deviance F-tests to evaluate goodness-of-fit

between a model including zone and all treatments to a zone-only model. For all Fucus

response metrics, the model including only zone was the most parsimonious model

Page 137: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

119

(Table B4.1). Hence, we used models including only zone for each response metric to test

the immersion stress hypothesis. To assess the competition and herbivory hypotheses, we

ran GLMs including zone and the set of treatments that tested each hypothesis. For the

competition hypothesis we compared canopy removal treatments to unmanipulated

controls. To test the herbivory hypothesis, we compared herbivore removals, paint

controls, and unmanipulated controls.

To assess the potential for recruit limitation or facilitation by barnacles we

assessed the relationship between Fucus recruit density and the percent cover of

barnacles, adult Fucus, and a barnacle:adult Fucus interaction. For this we used

generalized linear models with the quasipoisson distribution and log link function, to

account for overdispersion.

We used analysis of variance for statistical analysis of lab mesocosm results. For

each Fucus response metric (growth in length, growth in weight, new tip growth, new

receptacle formation, final receptacle proportion, elemental content) we tested the

influence of immersion time, light level, and the interaction between the two factors.

RESULTS

Field tests of vertical limitation

We observed striking differences in the condition of Fucus individuals

transplanted to the low zone compared to high zone transplants and marked individuals

over the course of the three-month experiment (Fig. 4.1). Low zone transplants lost much

of their thallus tissue on either side of the midrib. This observational difference in

condition was confirmed statistically as all Fucus metrics from the field experiment were

Page 138: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

120

greater in high zone individuals except for the final receptacle, and the percent nitrogen

content (%N) (Table 4.1), neither of which were related to zone.

Field experiments revealed that Fucus growth (in length) was less in the low zone

compared to the high zone (Fig. 4.2a; GLM, t56= -4.167, P= 0.0001). Additionally, high

zone transplants had greater new tip growth than low zone transplants (Fig. 4.2b; GLM,

t56= -5.053, P= 0.0001). High zone Fucus individuals formed more receptacles than low

zone individuals between the start and end of the experiment (Fig. 4.3a; GLM, t56= -

2.412, P= 0.019). However, there was no difference between zones in the final receptacle

proportion (Fig. 4.3b, GLM, t56= -1.105, P= 0.269).

C:N and carbon content (%C) of Fucus transplants were both lower in low zone

individuals compared to those in the high zone. Percent carbon in low zone individuals

was lower than that of high zone individuals (Fig. 4.4a; GLM, t21= -3.28, P= 0.004), and

low zone Fucus C:N was lower than high zone Fucus (Fig. 4.4c; GLM, t21= -2.32, P=

0.03). In contrast to C:N and %C, Fucus nitrogen content (%N) did not differ between

zones (Fig. 4.4b; GLM, t22= -0.919, P= 0.368).

To test the canopy competition hypothesis we compared canopy removal to

unmanipulated control treatments. Removal of the macroalgal canopy did not have an

effect on the growth in length, and there was no effect of zone (Fig. 4.5a, GLM, treatment

F1,1= 2.065, P= 0.171, zone F1,1= 3.499, P= 0.081, treatment:zone F1,1= 0.393, P= 0.540).

New tip growth did not differ between canopy treatments, but was greater in the high

zone (Fig. 4.5b, GLM, treatment F1,1= 0.836, P= 0.375, zone F1,1= 10.334, P= 0.006,

treatment:zone F1,1= 0.320, P= 0.580). Similar to growth in length, new receptacle

Page 139: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

121

formation did not differ between treatments or zone (Fig. 4.5c, GLM, treatment F1,1=

0.147, P= 0.707, zone F1,1= 3.885, P= 0.067, treatment:zone F1,1= 0.095, P= 0.762).

To test the herbivory hypothesis, we compared three treatments: herbivore

removal, paint control, and unmanipulated control. When accounting for herbivory

treatment, there was no effect of herbivory on Fucus transplant growth in length or in

new tip growth, but growth was greater in the high zone (Fig. 4.6; GLM; length:

treatment F1,2= 0.6587, P= 0.526, zone F1,1= 11.134, P= 0.002, treatment:zone F1,2=

0.535, P= 0.592; tips: treatment F1,2= 2.3877, P= 0.111, zone F1,1= 38.056, P< 0.0001,

treatment:zone F1,2= 1.817, P= 0.182). Fucus new receptacle formation was greater in the

high zone, but, like the growth metrics, there was no effect of treatment (GLM, treatment

F1,2= 1.281, P= 0.294, zone F1,1= 7.281, P= 0.012, treatment:zone F1,2= 1.060, P= 0.360).

For both metrics of growth (length and new tip growth) and reproductive status,

there was no difference between marked individuals and high zone transplants, indicating

no effect of the transplantation method itself (growth in length: t30= 0.854, P= 0.4; new

tip growth: t30= 0.501, P= 0.620; new receptacle formation: t30= 0.067, P= 0.947; final

receptacle proportion: z= 0.03, P= 0.976).

Fucus recruit densities were positively associated with primary cover of small

barnacles (Fig. B4.1, t79= 2.213, P= .0299), but not with adult Fucus cover (GLM, t78=

1.511, P= 0.135). There were no Fucus recruits found below the mid zone.

Mesocosm tests of vertical limitation

In lab mesocosms, Fucus growth (length) was affected by light, but not

immersion time or an interaction between the two factors (Fig. 4.7a, b; ANOVA,

Page 140: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

122

immersion: F1,53= 0.183, P= 0.67, light: F1,53= 5.851, P= 0.019, immersion:light: F1,53=

0.487, P= 0.488). Fucus growth in wet weight followed the same patterns (Fig. 4.7c, d;

ANOVA, immersion: F1,53= 1.753, P= 0.191, light: F1,53= 5.201, P= 0.027,

immersion:light: F1,53= 0.036, P= 0.850). Specifically, shaded Fucus individuals grew

0.17cm less in length, on average, than those in full light (± -0.457, 0.116, 95% CI) and

gained on average 0.487g less than those in full light (± -1.156, 0.182, 95% CI). Fucus

new tip growth was not related to immersion time or light in the mesocosm experiment

(Fig. 4.7e, f; ANOVA, immersion: F1,53= 3.118, P= 0.083, light: F1,53= 0.170, P= 0.681,

immersion:light: F1,53= 1.418, P= 0.239).

The presence of receptacles on Fucus individuals was decreased in both the high

immersion time (= low zone) treatments (Fig. 4.7g; ANOVA, F1,53= 8.787, P= 0.004) and

the shaded treatments (Fig. 4.7h; ANOVA, F1,53= 6.525, P= 0.013; no interaction;

ANOVA, F1,53= 0.378, P= 0.541).

Physiological measurements

Of the four photosynthetic parameters, none varied between immersion time or

light treatments, and there were no interactions (Table 4.2). Photosynthesis versus

irradiance curves fit to measurements of net photosynthesis at each irradiance overlapped

between light treatments (Fig. 4.8).

Elemental analysis for carbon and nitrogen content did not reveal an effect of

treatment on elemental content. Specifically, for carbon, there was no difference in %C

between immersion time (Fig. 4.9a; F1,16= 0.79, P= 0.39), light levels (Fig. 4.9b; F1,16=

1.98, P= 0.18), or an interactive effect (F1,16= 0.013, P= 0.91). Likewise, nitrogen content

Page 141: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

123

did not vary for any treatments (Fig. 4.9c, d; Immersion: F1,16= 0.003, P= 0.96; Light:

F1,16= 0.2.06, P= 0.17; Immersion:Light: F1,16= 1.30, P= 0.27). Carbon to nitrogen ratios

did not differ between immersion times (Fig. 4.9e; F1,16= 0.009, P= 0.92), light levels

(Fig. 4.9f; F1,12= 0.1.60, P= 0.22), nor was there an interaction between treatments (F1,12=

1.57, P= 0.23).

DISCUSSION

Our results suggest that light limitation is a possible mechanism behind poor

performance of adult Fucus in the low intertidal zone, contributing to setting its lower

distribution limit. Across both experiments, two primary findings are evident: 1) in the

field, Fucus grew more in the high zone compared to the low zone, irrespective of the

presence of grazers or the neighboring canopy, and exhibited overall symptoms of poor

condition in the low zone; and 2) in the mesocosms, Fucus had greater growth in the full

light treatments compared to the shade, regardless of immersion time. Although we

expected that Fucus would fall into the dominant paradigm of intertidal zonation (i.e.,

physically limited upper limit and biologically limited lower zone), results from the field

experiment were consistent with our hypothesis that a physical mechanism (immersion-

related stress) sets the lower limit for this species (H1). Mesocosm experiment results

suggested that the mechanism underlying this was not the greater immersion time in the

low zone as expected, but lower light conditions in that zone. Indeed, Fucus growth in

length, weight, and tips in the mesocosm experiment, was even slightly higher in the

immersion time treatment mimicking low zone conditions, suggesting that “drowning” is

not a viable mechanism in setting this species’ lower limit.

Page 142: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

124

Previous observations on this species have proposed that immersion time drives

vertical limits for Fucus, and in particular that the poor low zone performance of Fucus

was due to greater submergence in that zone (Druehl and Green 1982, Wright et al. 2004,

Dethier and Williams 2009). Our results suggest that it is not submergence or “drowning”

per se that limits Fucus distributions, but that in the low zone, lower light levels

associated with greater submergence may lead to poor condition, and possibly even

mortality, of Fucus. A light limitation mechanism for the hypothesis that immersion time

sets Fucus’ lower limits was proposed by Dethier and Williams (2009), but until now has

not been experimentally tested.

Yet another alternative, that fungal infection may play a role in setting the lower

limit for high zone seaweeds such as Fucus has been proposed (Zuccaro et al. 2008).

Evidence for this hypothesis is purely anecdotal, however, and further pursuit of this

alternative would be warranted.

Although there are many studies on photosynthesis in intertidal seaweeds, the role

of the intertidal light environment is often overlooked and evidence for the role of light in

setting intertidal species distributions is limited. In part, this is because photosynthesis

can vary between periods of emersion and immersion not only due to light availability,

but also as a function of temperature and the rate of desiccation during emersion.

Research on the variation in photosynthetic rate during immersion and emersion shows

mixed results. Some studies indicate that, barring desiccation, rates of photosynthesis in

Fucus tend to be higher during emersion (Dring and Brown 1982, Williams and Dethier

2005). But Williams and Dethier (2005) found that photosynthesis in water was

Page 143: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

125

consistently much higher than in air. Madsen and Maberly (1990) found similar results,

but also detected an interactive effect of temperature, wherein photosynthesis was greater

in air at higher temperatures, but greater in water under colder conditions. In our field

experiment, carbon content was greater in the high zone, consistent with the proposed

mechanism of light limitation in the low zone. This does suggest a difference in C-

fixation potentially associated with the photosynthetic process. We did not see any

corresponding difference in carbon content or photosynthetic parameters between

mesocosm light or immersion time treatments.

It is possible that the overall lower light levels in the mesocosm experiment

compared to levels in the field contributed to our results. While relative light levels in the

full and shaded light treatments were different, the absolute levels in both treatments

were lower than light conditions would be in a field setting. This may have stressed

individuals in all treatments, possibly accounting for the absence of a pattern in %C or

photosynthetic parameters as well. Additionally, the photosynthetic measurements were

taken after 30-32 days in the experimental treatments, perhaps too short in duration to see

a response in the photosynthetic apparatus to laboratory conditions.

New tip growth did not differ between mesocosm light or immersion time

treatments, though new tip growth was low for both treatments (<1 new tip per individual

on average). In Fucus, new tips are produced as fronds lengthen and branch. By our field

convention, we counted new tips only if they exceeded 1cm from the last dichotomy.

However, Fucus growth rate in the mesocosm experiment was not great enough to result

Page 144: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

126

in enough new tips >1cm. Thus, the absence of an effect of light on new tip growth likely

stems from low statistical power and differences between field and laboratory conditions.

Effects on life history and elemental content

Low zone conditions evidently had a negative effect on Fucus reproductive output

as well as growth. In the field experiment, reproductive output was greater in high zone

individuals, and in the mesocosm experiment reproductive output was greater in both the

low immersion time and full light treatments (the mesocosm analogs for high zone

conditions). New receptacle formation was the only measurement for which there was a

difference between immersion time treatments in the mesocosm experiment.

Reproductive output may be particularly sensitive to immersion time, as it can directly

impact success of gametes in fertilization and settling (Lobban and Harrison 1994). Thus,

we would expect immersion time to have a greater effect on reproductive output than on

the other metrics used.

Nitrogen content did not differ between zones in field or mesocosm experiment

individuals, counter to our expectations. Since seaweeds assimilate nutrients only while

submerged (Lobban and Harrison 1994), we expected content to differ due to the vastly

different immersion times in the two zones. However, nitrate uptake is known to be

higher in high zone species, especially after periods of desiccation (Thomas et al. 1987,

Bracken et al. 2011). Fucus adaptations to high zone conditions may enable them to take

up nutrients more quickly in the high zone (S. Close, unpublished data) to compensate for

reduced immersion times there, thus resulting in equal nitrogen content across zones.

Page 145: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

127

Role of biotic interactions and recruitment in Fucus distribution

As predicted by theory (Connell 1972, Menge and Sutherland 1987, Bruno et al.

2003), we expected that species interactions would play some role in setting the lower

limits of Fucus’ distribution. However, we did not observe an effect of canopy

competition (H2) or herbivory (H3) on Fucus transplants in either zone. Although low

zone Fucus transplants tended to exhibit greater new tip growth in canopy removal

treatments (implying some potentially weakly inhibitory effect of the macroalgal canopy

on adult Fucus growth), an opposite trend in growth occurred (Fig. 4.5).

The paucity of herbivores in the low zone compared to the high zone likely played

a role in the lack of an herbivore effect on Fucus growth since the difference in herbivore

density between control and removal treatments was much smaller. Although some

herbivores are known to cross a copper paint barrier (e.g. the gastropod Chlorostoma

funebralis, the sea urchin Strongylocentrotus purpuratus), we observed very few

individuals of those species in our plots. Further, S. purpuratus is sedentary, rarely

moving out of the depressions they carve into the rock, and our experiments were placed

in areas away from sea urchin beds. Additionally, there was no overlap in herbivore

species between zones (S. Close, personal observations), so any herbivory on Fucus in

the low zone would represent a novel interaction. However, Thornber and colleagues

(2008) found in cross-zone feeding trials that herbivores tended to prefer algae from the

zone opposite to their own, so it is unlikely that the lack of an effect of herbivores was

due to Fucus being a novel food source in the low zone. Fucus is known to be heavily

defended by phenolic compounds that deter herbivores, possibly also contributing to the

Page 146: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

128

absence of an herbivore effect (Van Alstyne 1988). Thus, the poor condition of Fucus

transplants in the low zone was independent of the presence or absence of grazers or a

macroalgal canopy, suggesting causation by a physical mechanism unrelated to species

interactions.

Our observations of the distribution of Fucus recruits and primary cover of adult

Fucus and barnacles indicated that dispersal distance and barnacle facilitation of Fucus

recruitment may contribute to determining Fucus’ intertidal distribution. Fucus recruits

were not found in the field observations below the mid-high zone, and were most

plentiful in the high zone, lending support to H4 that limited dispersal distances may help

constrain Fucus’ intertidal limits, as seen in other studies on this species and other

intertidal macroalgae species (van Tamelen et al. 1997, Shanks et al. 2003). High

densities of Fucus recruits co-occurred with high primary cover of barnacles (H5), but we

did not quantify the number of recruits that were attached to barnacles. Barnacles are

known to facilitate the recruitment of other intertidal species (Farrell 1991), and may

provide enhanced substrate rugosity that Fucus recruits require to attach (Lobban and

Harrison 1994). In a study done in Alaska, Fucus disproportionately recruited to

barnacles, but those individuals in the study that were attached to barnacles became

dislodged with much less force than those attached to bare rock (van Tamelen and Stekoll

1997). At our study site, adult Fucus are not found attached to barnacles at a

disproportionately high level but it does occur commonly (S. Close, personal

observations). Experimental evaluation of these hypotheses (H4 and H5), as well as

Page 147: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

129

investigating H1 through H3 across other life stages of Fucus, are important future

avenues of research.

Responses to environmental gradients

It is important to consider the capacity for phenotypic plasticity and/or local

adaptation to influence species responses to environmental gradients. Given the limited

dispersal distances of many intertidal macroalgal species, there is the potential for local

adaptation to arise under certain conditions. For instance, Hays (2007) found evidence for

local adaptation, but not phenotypic plasticity, in the high intertidal fucoid alga Silvetia

compressa. Interestingly, though, the presence of local adaptation in these populations

depended on the steepness of the environmental gradients, and so varied on a site by site

basis. Similar patterns in local adaptation are plausible for Fucus. Variation among sites

in Fucus morphology and distribution was evident along the Oregon coast (S. Close,

personal observations), so understanding the capacity for local adaptation along

environmental gradients in this species is important. Further, the question of what

constitutes stressful conditions can be defined very differently both among and within

species, and, for Fucus, has been shown to vary seasonally as well (Dethier and Williams

2009).

Environmental stress models predict that abiotic factors are important to

community structure at high levels of environmental stress, but at more benign conditions

species interactions predominate. Our results are consistent with some aspects of ESMs,

in that abiotic factors are important in areas of high stress, but indicate that for this

species the low zone is in fact more stressful, contradicting the classic paradigm of rocky

Page 148: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

130

intertidal zonation. Intertidal species’ boundaries result not just from simple spatial or

size refugia, but from complex equilibria encompassing such factors as predation or

consumption, population gain (growth or recruitment), and spatial gradients in these

factors (Menge and Sutherland 1987, Bruno et al. 2003). Our findings are consistent with

this concept. We determined that a physiological mechanism inhibits Fucus growth in the

lower intertidal, and found compelling evidence that light, potentially among other

factors, plays a key role. But recruitment and facilitation of recruitment by barnacles also

may influence the distributional capacity for Fucus in the intertidal. The interaction of

these factors, potentially among others such as local population dynamics and

environmental gradients, determines the realized boundaries of Fucus.

Extending these observations to a broader gradient of conditions in the rocky

intertidal is important in determining the generality or context dependency of this work.

Here we only tested the lower limit of adult Fucus, but understanding what drives upper

limits or distributions across gradients is an important question that has implications to

potential distribution shifts under climate change. There are few documented examples of

biotic interactions setting the upper limits for intertidal species (but see Cubit 1984), so

we expect that Fucus’ upper limits would be set by abiotic factors such as desiccation or

heat stress. This may make Fucus an unusual case in that both its upper and lower limits

may be set by abiotic factors. Further, known differences in light regime exist along this

coast providing an ideal setting for a comparative experimental approach to testing the

generality of our findings. Coastal oceanography, specifically the process of upwelling,

and coastal bathymetry interact to modify the light environment, primarily via shading by

Page 149: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

131

phytoplankton blooms, along the Oregon coast (Menge et al. In Prep., Kavanaugh et al.

2009). Light availability in rocky intertidal systems has the capacity to influence benthic

macrophyte distributions (this study), growth (Kavanaugh et al. 2009), life history

strategies (Dethier et al. 2005), and influence community structure (Harley 2002). We are

just starting to understand the role light plays in these communities, and here we have

shown that it can have important consequences for species distributions and growth.

Through determining a physical mechanism behind observed distribution patterns

in a common intertidal alga, we propose that Fucus is an example of a species for which

both the upper and lower limits are set by abiotic factors. Thus, the paradigm that lower

limits of intertidal species distributions are set by biotic interactions may not always hold.

As this paradigm is frequently the basis for predictions of the impacts of sea level rise on

intertidal communities, our findings cast light on the complexities of intertidal zonation

and the difficult task ahead of predicting climate change impacts in this system.

ACKNOWLEDGEMENTS

We wish to thank a number of people who assisted with the field portion of this

study: E. Cerny-Chipman, C. Shen, M. Prechtl; and the mesocosm portion of the study:

A. Iles, the Hatfield Marine Science Center facilities staff (especially J. Lawson and T.

Hardman), M. Atkinson, M. Beazley, K. Krallman, and J. Robertson. We are also grateful

to C. Blanchette, for advice on transplant methods, E. Carrington, for advice on

mesocosm setup, and L. Tait, for assistance with photosynthesis measurements and

calculations. Comments and feedback from A. Iles, B. Beechler, E. Gorsich, and others

greatly improved this manuscript. This research was supported by a National Science

Page 150: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

132

Foundation Graduate Research Fellowship, the Hatfield Marine Science Center Mamie

Markham Research Award, Department of Zoology Research Funds (all to SLC), by an

endowment from the Wayne and Gladys Valley Foundation, and by grants from the

David and Lucile Packard and Gordon and Betty Moore Foundations, which supported

PISCO, the Partnership for Interdisciplinary Studies of Coastal Oceans.

Page 151: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

133

FIGURES

Figure 4.1. Photos of transplanted Fucus individuals at the end of the three-month field experiment. (a) Naturally-growing Fucus;

(b) high zone transplants; (c) low zone transplants. Note that plant in (c) looks substantially less healthy than those in (a) or (b).

Both transplants shown here were in unmanipulated control treatments.

a b c

Page 152: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

134

Figure 4.2. Growth of Fucus transplants to the high and low intertidal zones, as measured by length (a) and new tip growth (b).

Different letters indicate statistically significant differences between zones at the alpha= 0.05 level. Bars show means ± SE.

Page 153: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

135

Figure 4.3. New receptacle formation (a) and final receptacle proportion (b) in the high and low zone Fucus transplants. See Fig.

4.2 for further explanation.

Page 154: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

136

Figure 4.4. Elemental content of Fucus transplants in each zone (a) %C, b) %N, c) C:N). See Fig. 4.2 for further explanation.

Page 155: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

137

Figure 4.5. Growth in length (a) and tips (b) and new receptacle formation (c) of transplants of both zones (black bars= high zone,

white bars= low zone) and the two treatments to test the canopy competition hypothesis. Bars on the left show mean growth (±

SE) in canopy removal treatments and bars on the right show mean growth (± SE) in unmanipulated control transplants. See Fig.

4.2 for further explanation.

Page 156: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

138

Figure 4.6. Growth in length (a) and tips (b) and new receptacle formation (c) of transplants of both zones (black bars= high zone,

white bars= low zone) and the three treatments testing the hypothesis that herbivory sets the lower limit in Fucus. Bars on the left

show unmanipulated control transplants, middle bars show paint controls, and the bars on the right show herbivore removal

treatments. See Fig. 4.2 for further explanation.

Page 157: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

139

Figure 4.7. Mesocosm results showing the influence on Fucus of immersion time and

light level treatments. Zone code: high zone = low immersion time; low zone = high

immersion time. Light code: full = no cover, shade = mesh cover. Plots show Fucus

growth (change in length; a), weight (b), change in number of tips (c), and reproductive

status (d; change in number of receptacles from day 1 to day 32). See Fig. 4.2 for further

explanation.

Page 158: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

140

Figure 4.8. Net photosynthesis (g O2 g-1

h-1

) across irradiances (µmol photons m-2

s-1

) for

Fucus individuals in full (filled circles) and shade (open circles) light treatments. Curves

show Michaelis-Menten fits based on photosynthetic parameters calculated from a linear

fit of net photosynthesis measured at 0 and 153 µmol photons m-2

s-1

. Bars show mean (±

SE).

Page 159: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

141

Figure 4.9. Elemental content (%C, %N, and C:N) of Fucus individuals from the

mesocosm experiment. Shown are the results from immersion time (a, c, e) and light

level (b, d, f) treatments. Zone code: high zone = low immersion time; low zone = high

immersion time. Light code: full = no cover, shade = mesh cover. Rows show Fucus

growth (change in length; a, b), weight (c, d), change in number of tips (e, f), and

reproductive status (g, h; change in number of receptacles from day 1 to day 32). See Fig.

2 for further explanation.

Page 160: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

142

TABLES

Table 4.1. Results from generalized linear models relating each metric of Fucus growth,

reproductive status, and elemental content to zone. Alpha= 0.05, significant p-values are

bold. All models: Response = β0 + β1zone + ε.

Response variable Estimate (± SE) P-value Effect direction

change in length (cm) - 2.867 (0.688) 0.0001 greater in high zone

change in # tips (#) - 18.555 (3.672) 0.0001 greater in high zone

change in # receptacles (#) -3.142 (1.303) 0.019 greater in high zone

final receptacle proportion -1.087 (0.984) 0.269 no effect

% Carbon -5.006 (1.528) 0.03 greater in high zone

% Nitrogen -0.060 (0.065) 0.368 no effect

Carbon:Nitrogen -2.352 (1.014) 0.004 greater in high zone

Page 161: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

143

Table 4.2. Results of two-way ANOVAs testing the influence of both factors, and their

interaction, on each photosynthetic parameter: Pmax, Respiration, Slope, and

Compensating Irradiance. Parameters were calculated from a linear fit between the rates

measured at light levels 0 (respiration) and 153 µmol photons m-2

s-1

.

Pmax Respiration Slope Ec

(g O2 gDW-1

h-1

) (g O2 gDW-1

h-1

) (alpha) µmol photons m-2

s-1

P= 0.907 0.286 0.864 0.495

F1,6= 0.015 1.374 0.032 0.527

P= 0.745 0.9939 0.433 0.869

F1,6= 0.116 0.0001 0.706 0.03

P= 0.769 0.901 0.784 0.335

F1,6= 0.095 0.017 0.082 1.099

Zone (immersion time)

Light

Zone:Light

Page 162: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

144

CHAPTER 5 – GENERAL CONCLUSION

The research presented here illustrates the role of environmental variation,

particularly but not exclusively in terms of nutrient dynamics, in macrophyte

assemblages. Regulation of communities through bottom-up inputs is known to play a

role in many communities including nearshore marine systems (Wallace et al. 1999,

Carpenter et al. 2001, Worm et al. 2002, Menge et al. 2004, Menge and Menge 2013), but

our understanding of how this influences macrophytes is limited. By approaching this

issue from different scales and perspectives, I have gained valuable insights into how the

physical environment affects this productive, diverse community.

By employing a large-scale, long-term approach in Chapters 2 and 3, I assessed

the role of nutrient availability in macrophyte ecology at spatiotemporal scales that

allowed evaluation of the role of upwelling and climate-scale oceanographic patterns. In

Chapter 2, I investigated the drivers of tissue nutrient content (proxied by carbon to

nitrogen ratios, C:N) in four dominant intertidal macrophyte species: Fucus distichus,

Mazzaella splendens, Phyllospadix scouleri, and Saccharina sessilis. Tissue nutrient

content in macrophytes can be influenced by both environmental and physiological

factors (Lobban and Harrison 1994), making variation in this measure complex and

difficult to tease apart. My analysis focused on understanding the spatiotemporal

variation in C:N, and the environmental drivers (including nutrient availability,

upwelling, and basin-scale climate indices) underlying that variation. I found that the four

species studied all varied in not only their C:N, but how that ratio was related to nutrient

availability and other environmental drivers. In all species but F. distichus C:N was

Page 163: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

145

negatively related to nitrogen availability, but the nature of that relationship varied across

levels of nitrogen availability. Generally, patterns of spatiotemporal variability as well as

analyses on environmental drivers indicated that upwelling influences C:N to some extent

in all species. Consistency in interannual variability across species and space further

suggested that oceanographic variables drive macrophyte nutrient dynamics.

In addition to nutrient content, assessing the influence of nutrient availability on

growth and how that interacts with nutrient content can provide new insights into the role

of nutrient dynamics in macrophyte assemblages. In Chapter 3, I considered the roles of

two potential limiting factors of growth in S. sessilis and P. scouleri: nutrient and light

availability. I found indications that their growth responded to nutrient and light

availability differently, but evidence for a relationship between growth and C:N was

limited. Specifically, my results suggested that growth in S. sessilis was both positively

associated with nutrient availability, and negatively associated with light availability.

While seemingly contradictory, this is consistent with previous research (Kavanaugh et

al. 2009), as well as with our understanding of the interaction between nutrient and light

availability that occurs during upwelling events and is associated with the formation of

phytoplankton blooms (Menge and Menge 2013). On the other hand, P. scouleri growth

showed complex patterns in how it related to light and nutrient availability. For both

resources, the response of growth depended on the level of the resource. For nutrients,

growth was enhanced at low to intermediate nutrient concentrations, but declined when

nutrient concentrations were high. This is consistent with nutrient toxicity which is

known to occur in seagrasses, and with my findings in Chapter 2 that C:N in P. scouleri

Page 164: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

146

was particularly sensitive to environmental variation and appeared negatively impacted at

high nutrient concentrations. Similar patterns were evident for light availability and P.

scouleri growth. So for Chapter 3, I again found that while the relationship of

macrophytes to nutrient availability is complicated, upwelling dynamics underlie

observed patterns in macrophyte growth.

Results from Chapter 2 suggested that nutrient content was low (C:N was high) in

F. distichus, leading me to speculate on the potential reasons for the difference. F.

distichus lives in the high intertidal zone where it is frequently emersed and unable to

take up nutrients for long periods of time relative to the low zone (Lobban and Harrison

1994). I surmised that nutrients may be a factor in the vertical distribution of this species,

and designed a field experiment to test the factors that set its lower limits. My expectation

was that F. distichus would adhere to the classic model of rocky intertidal zonation:

upper limits are set by abiotic factors and lower limits are set by biotic interactions. I

found that irrespective of the presence of grazers or a macrophyte canopy, F. distichus

did poorly in the low zone, suggestive of abiotic factors setting its lower limit. I used a

laboratory mesocosm experiment to test the potential mechanism(s) underlying poor

performance of F. distichus in that zone, and focused on immersion time and light level.

My results suggested a role of light availability in setting the lower limit of F. distichus.

Additionally, I demonstrated that this species deviates from the classic model of intertidal

zonation. This model is often the basis for predictions of the impacts of climate change,

specifically sea level rise, on intertidal communities. Our findings indicate that this

Page 165: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

147

species may respond to sea level rise differently than predicted for other intertidal

organisms.

By pairing my large-scale observations with this local-scale experimental

approach, I was able to generate hypotheses and test them in the field and lab. The

combined results from Chapters 2 and 4 gave me insights into how F. distichus is

affected by environmental variability. Nutrient availability and C:N were not related in F.

distichus, nor did we see evidence that nutrient availability was a factor in zonation of F.

distichus. Paradoxically, analyses in Chapter 2 suggested that basin- and region-scale

patterns did have an impact on the nutrient status of this species. Reconciling these

results, it seems that F. distichus could integrate the signal of nutrient variability over

longer periods of time and may not be as sensitive to high-frequency variation in the

nutrient environment in this system. The implications of this for bottom-up regulation of

this species, as well as for sensitivity to environmental perturbations, should be

considered.

From the research presented here our understanding of how intertidal

macrophytes are influenced by their physical environment, including but not limited to

nutrient dynamics, has been advanced. In Chapters 2 and 3 I found that even in this

nutrient-replete ecosystem, nutrient availability influences macrophytes both in terms of

their tissue nutrient content and growth. The relation of both of these measures to

community dynamics suggests possible mechanisms through which nutrient availability

could scale up to influence bottom-up regulation of intertidal communities.

Page 166: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

148

Relatedly, a second important insight gained here is that species identity matters

to the role of nutrient dynamics in macrophyte assemblages. My investigations focused

on four taxonomically diverse species, and for no two species were the responses to

nutrient dynamics or environmental variability the same. While I did observe some

consistencies, such as in spatial variation in C:N between M. splendens and S. sessilis,

more often than not my results depend on species identity. In Chapter 2, however, I saw

that among-species variation in C:N generally fell out as predicted based on species traits.

Understanding which traits are important to these processes is an important step towards

improving our abilities to predict community-level responses to environmental change.

A third insight gained from this research is that light availability can influence

growth and intertidal distributions. Light availability is emerging as an important

structuring agent in intertidal communities, although we know very little about how it

influences these systems (but see: Harley 2002, Kavanaugh et al. 2009). In the CCLME,

variations in light availability arise primarily due to the downstream consequences of

upwelling. As upwelling draws nutrient-rich waters to the ocean’s surface, large

phytoplankton blooms form which attenuate light, particularly in places where the

continental shelf is wide and retention is high (Chan et al. 2008). But, it is important to

consider that light availability can be high both during periods of high nutrient

availability and low nutrient availability, such as downwelling. That variations in light

arise due to upwelling suggests that this process, and the processes it influences, may be

sensitive to global change.

Page 167: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

149

The oceans and the processes that shape them are changing in very fundamental

ways. Upwelling is expected to intensify under global climate change scenarios (Bakun

1990), and in the California Current Large Marine Ecosystem (CCLME) this has already

been documented. Upwelling events are occurring less frequently, and when they do

occur are stronger and longer (Iles et al. 2012). In this body of work, I found that

environmental variation related to upwelling consistently plays a central role in

structuring intertidal macrophyte assemblages. Thus, the ramifications to changes in

upwelling in the CCLME are potentially immense and underscore the critical need to

better understand how environmental variation influences macrophyte assemblages.

Page 168: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

150

BIBLIOGRAPHY

Abbott, I. A., and G. J. Hollenberg. 1976. Marine algae of California. Stanford University

Press, Stanford Calif.

Ågren, G. I. 2008. Stoichiometry and Nutrition of Plant Growth in Natural Communities.

Annual Review of Ecology, Evolution, and Systematics 39:153–170.

Van Alstyne, K. 1988. Herbivore grazing increases polyphenolic defenses in the intertidal

brown alga Fucus distichus. Ecology 69:655–663.

Ang, P. 1991. Age-Dependent and Size-Dependent Growth and Mortality in a Population

of Fucus distichus. Marine Ecology Progress Series 78:173–187.

Aquilino, K. M., M. E. S. Bracken, M. N. Faubel, and J. J. Stachowicz. 2009. Local-scale

nutrient regeneration facilitates seaweed growth on wave-exposed rocky shores in

an upwelling system. Limnology and Oceanography 54:309–317.

Atkinson, M. J., and S. V. Smith. 1983. C:N:P Ratios of Benthic Marine Plants.

Limnology and Oceanography 28:568–574.

Baker, S. M. 1909. On the causes of the zoning of brown seaweeds on the seashore. New

Phytologist 8:196–202.

Bakun, A. 1967. Daily and weekly upwelling indices, west coast of North America. Page

114. United States Department of Commerce, National Oceanic and Atmospheric

Administration, Seattle, WA.

Bakun, A. 1990. Global Climate Change and Intensification of Coastal Ocean Upwelling.

Science 247:198–201.

Barth, J. A., B. A. Menge, J. Lubchenco, F. Chan, J. M. Bane, A. R. Kirincich, M. A.

McManus, K. J. Nielsen, S. D. Pierce, and L. Washburn. 2007. Delayed upwelling

alters nearshore coastal ocean ecosystems in the northern California current.

Proceedings of the National Academy of Sciences 104:3719 –3724.

Barth, J. A., S. D. Pierce, and R. L. Smith. 2000. A separating coastal upwelling jet at

Cape Blanco, Oregon and its connection to the California Current System. Deep

Sea Research Part II: Topical Studies in Oceanography 47:783–810.

Bates, D., M. Maechler, and B. Bolker. 2013. lme4: Linear mixed-effects models using

S4 classes.

Benedetti-Cecchi, L., and F. Cinelli. 1997. Confounding in field experiments: direct and

indirect effects of artifacts due to the manipulation of limpets and macroalgae.

Journal of Experimental Marine Biology and Ecology 209:171–184.

Bjorkstedt, E. P., R. Goericke, S. McClatchie, E. Weber, W. Watson, N. Lo, B. Peterson,

B. Emmett, R. Brodeur, J. Peterson, M. Litz, J. Gomez-Valdez, G. Gaxiola-

Castro, B. Lavaniegos, F. Chavez, C. A. Collins, J. Field, K. Sakuma, P.

Warzybok, R. Bradley, J. Jahncke, S. Bograd, F. Schwing, G. S. Campbell, J.

Hildebrand, W. Sydeman, S. A. Thompson, J. L. Largier, C. Halle, S. Y. Kim, and

J. Abell. 2011. State of the California Current 2010-2011: Regionally Variable

Responses. California Cooperative Oceanic Fisheries Investigations Reports

52:36–68.

Page 169: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

151

Bjorkstedt, E. P., R. Goericke, S. McClatchie, E. Weber, W. Watson, N. Lo, B. Peterson,

B. Emmett, J. Peterson, R. Durazo, G. Gaxiola-Castro, F. Chavez, J. T.

Pennington, C. A. Collins, J. Field, S. Ralston, K. Sakuma, S. J. Bograd, F. B.

Schwing, Y. Xue, W. J. Sydeman, S. A. Thompson, J. A. Santora, J. Largier, C.

Halle, S. Morgan, S. Y. Kim, K. P. B. Merkens, J. A. Hildebrand, and L. M.

Munger. 2010. State of the California Current 2009-2010: Regional Variation

Persists Through Transition from La Nina to El Nino (and Back?). California

Cooperative Oceanic Fisheries Investigations Reports 51:39–69.

Blanchette, C. A. 1997. Size and survival of intertidal plants in response to wave action: a

case study with Fucus gardneri. Ecology 78:1563–1578.

Blanchette, C. A., B. R. Broitman, and S. D. Gaines. 2006. Intertidal community structure

and oceanographic patterns around Santa Cruz Island, CA, USA. Marine Biology

149:689–701.

Blanchette, C. A., E. A. Wieters, B. R. Broitman, B. P. Kinlan, and D. R. Schiel. 2009.

Trophic structure and diversity in rocky intertidal upwelling ecosystems: A

comparison of community patterns across California, Chile, South Africa and

New Zealand. Progress in Oceanography 83:107–116.

Bracken, M. E. S., C. A. Gonzalez-Dorantes, and J. J. Stachowicz. 2007. Whole-

community mutualism: Associated invertebrates facilitate a dominant habitat-

forming seaweed. Ecology 88:2211–2219.

Bracken, M. E. S., E. Jones, and S. L. Williams. 2011. Herbivores, tidal elevation, and

species richness simultaneously mediate nitrate uptake by seaweed assemblages.

Ecology 92:1083–1093.

Bracken, M. E. S., and J. J. Stachowicz. 2007. Top-down modification of bottom-up

processes: selective grazing reduces macroalgal nitrogen uptake. Marine Ecology

Progress Series 330:75–82.

Brauer, V. S., M. Stomp, and J. Huisman. 2012. The Nutrient-Load Hypothesis: Patterns

of Resource Limitation and Community Structure Driven by Competition for

Nutrients and Light. The American Naturalist 179:721–740.

Broitman, B. R., S. A. Navarrete, F. Smith, and S. D. Gaines. 2001. Geographic variation

of southeastern Pacific intertidal communities. Marine Ecology Progress Series

224:21–34.

Bruno, J., J. Stachowicz, and M. Bertness. 2003. Inclusion of facilitation into ecological

theory. Trends in Ecology & Evolution 18:119–125.

Bulleri, F., B. D. Russell, and S. D. Connell. 2012. Context-Dependency in the Effects of

Nutrient Loading and Consumers on the Availability of Space in Marine Rocky

Environments. PLoS ONE 7:e33825.

Burkepile, D. E., and M. E. Hay. 2006. Herbivore vs. nutrient control of marine primary

producers: Context-dependent effects. Ecology 87:3128–3139.

Cade, B. S., J. W. Terrell, and R. L. Schroeder. 1999. Estimating effects of limiting

factors with regression quantiles. Ecology 80:311–323.

Carpenter, S. R., J. J. Cole, J. R. Hodgson, J. F. Kitchell, M. L. Pace, D. Bade, K. L.

Cottingham, T. E. Essington, J. N. Houser, and D. E. Schindler. 2001. Trophic

Page 170: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

152

cascades, nutrients, and lake productivity: whole-lake experiments. Ecological

Monographs 71:163–186.

Castelao, R. M., and J. A. Barth. 2005. Coastal ocean response to summer upwelling

favorable winds in a region of alongshore bottom topography variations off

Oregon. Journal of Geophysical Research 110:C10S04.

Chan, F., J. A. Barth, J. Lubchenco, A. Kirincich, H. Weeks, W. T. Peterson, and B. A.

Menge. 2008. Emergence of Anoxia in the California Current Large Marine

Ecosystem. Science 319:920–920.

Chavez, F. P., J. Ryan, S. E. Lluch-Cota, and M. Ñ. C. 2003. From Anchovies to Sardines

and Back: Multidecadal Change in the Pacific Ocean. Science 299:217–221.

Christie, H., K. M. Norderhaug, and S. Fredriksen. 2009. Macrophytes as habitat for

fauna. Marine Ecology Progress Series 396:221–233.

Connell, J. H. 1961. The influence of interspecific competition and other factors on the

distribution of the barnacle Chthamalus stellatus. Ecology 42:710–723.

Connell, J. H. 1972. Community Interactions on Marine Rocky Intertidal Shores. Annual

Review of Ecology and Systematics 3:169–192.

Connolly, S. R., B. A. Menge, and J. Roughgarden. 2001. A latitudinal gradient in

recruitment of intertidal invertebrates in the Northeast Pacific Ocean. Ecology

82:1799–1813.

Cubit, J. D. 1984. Herbivory and the seasonal abundance of algae on a high intertidal

rocky shore. Ecology 65:1904–1917.

Dayton, P. K. 1975. Experimental Evaluation of Ecological Dominance in a Rocky

Intertidal Algal Community. Ecological Monographs 45:137–159.

Dethier, M. N., and S. L. Williams. 2009. Seasonal stresses shift optimal intertidal algal

habitats. Marine Biology 156:555–567.

Dethier, M., S. Williams, and A. Freeman. 2005. Seaweeds under stress: Manipulated

stress and herbivory affect critical life-history functions. Ecological Monographs

75:403–418.

Downing, J. A. 1997. Marine nitrogen: phosphorus stoichiometry and the global N:P

cycle. Biogeochemistry 37:237–252.

Dring, M. J., and F. A. Brown. 1982. Photosynthesis of intertidal brown algae during and

after periods of emersion: A renewed search for physiological causes of zonation.

Marine Ecology Progress Series 8:301–308.

Druehl, L. D., and J. M. Green. 1982. Vertical distribution of intertidal seaweeds as

related to patterns of submersion and emersion. Marine Ecology Progress Series

9:163–170.

Duarte, C. M. 1990. Seagrass Nutrient Content. Marine Ecology Progress Series 67:201–

207.

Duarte, C. M. 1992. Nutrient Concentration of Aquatic Plants: Patterns Across Species.

Limnology and Oceanography 37:882–889.

Dudgeon, S. R., R. S. Steneck, I. R. Davison, and R. L. Vadas. 1999. Coexistence of

similar species in a space-limited intertidal zone. Ecological Monographs 69:331–

352.

Page 171: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

153

Dugdale, R. C. 1967. Nutrient Limitation in the Sea: Dynamics, Identification, and

Significance. Limnology and Oceanography 12:685–695.

Durazo, R., C. A. Collins, D. K. Hyrenbach, F. B. Schwing, T. R. Baumgartner, J. García,

D. Loya, R. L. Smith, P. A. Wheeler, S. J. Bograd, A. Huyer, R. J. Lynn, and W.

J. Sydeman. 2001. The state of California current, 2000 – 2001 : a third straight

La Niña year.

Farrell, T. M. 1991. Models and mechanisms of succession - an example from a rocky

intertidal community. Ecological Monographs 61:95–113.

Fox, J., and S. Weisberg. 2011. An {R} Companion to Applied Regression, Second

Edition. Sage, Thousand Oaks, CA.

Francis, R. C., and S. R. Hare. 1994. Decadal-scale regime shifts in the large marine

ecosystems of the Northeast Pacific: a case for historical science. Fisheries

Oceanography 3:279–291.

Freidenburg, T. L., B. A. Menge, P. M. Halpin, M. Webster, and A. Sutton-Grier. 2007.

Cross-scale variation in top-down and bottom-up control of algal abundance.

Journal of Experimental Marine Biology and Ecology 347:8–29.

Fujita, R. M., P. A. Wheeler, and R. L. Edwards. 1989. Assessment of macroalgal

nitrogen limitation in a seasonal upwelling region. Marine Ecology Progress

Series 53:293–303.

Gabrielson, P. W., T. B. Widdowson, and S. C. Lindstrom. 2006. Keys to the seaweeds

and seagrasses of southeast Alaska, British Columbia, Washington, and Oregon.

PhycoID : May be purchased from Sandra Lindstrom, [Vancouver, B.C.].

Gibbs, R. E. 1902. Phyllospadix as a Beach-builder. American Naturalist 36:101–109.

Gorman, D., B. D. Russell, and S. D. Connell. 2009. Land-to-sea connectivity: linking

human-derived terrestrial subsidies to subtidal habitat change on open rocky

coasts. Ecological Applications 19:1114–1126.

Graham, W., J. Field, and D. Potts. 1992. Persistent Upwelling Shadows and Their

Influence on Zooplankton Distributions. Marine Biology 114:561–570.

Graham, W. M., and J. L. Largier. 1997. Upwelling shadows as nearshore retention sites:

the example of northern Monterey Bay. Continental Shelf Research 17:509–532.

Gravel, D., F. Guichard, M. Loreau, and N. Mouquet. 2010. Source and sink dynamics in

meta-ecosystems. Ecology 91:2172–2184.

Guerry, A. D. 2008. Interactive effects of grazing and enrichment on diversity;

conceptual implications of a rocky intertidal experiment. Oikos 117:1185–1196.

Guerry, A. D., B. A. Menge, and R. A. Dunmore. 2009. Effects of consumers and

enrichment on abundance and diversity of benthic algae in a rocky intertidal

community. Journal of Experimental Marine Biology and Ecology 369:155–164.

Guiry, M. D., and G. M. Guiry. 2010. Fucus distichus subsp. evanescens (C.Agardh)

H.T.Powell. World-wide electronic publication. http://www.algaebase.org.

Harley, C. D. G. 2002. Light availability indirectly limits herbivore growth and

abundance in a high rocky intertidal community during the winter. Limnology and

Oceanography 47:1217–1222.

Harley, C. D. G. 2003. Abiotic stress and herbivory interact to set range limits across a

two-dimensional stress gradient. Ecology 84:1477–1488.

Page 172: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

154

Harley, C. D. G., K. M. Anderson, K. W. Demes, J. P. Jorve, R. L. Kordas, T. A. Coyle,

and M. H. Graham. 2012. Effects of climate change on global seaweed

communities. Journal of Phycology 48:1064–1078.

Harley, C. D. G., A. R. Hughes, K. M. Hultgren, B. G. Miner, C. J. B. Sorte, C. S.

Thornber, L. F. Rodriguez, L. Tomanek, and S. L. Williams. 2006. The impacts of

climate change in coastal marine systems. Ecology Letters 9:228–241.

Harpole, W. S., J. T. Ngai, E. E. Cleland, E. W. Seabloom, E. T. Borer, M. E. S. Bracken,

J. J. Elser, D. S. Gruner, H. Hillebrand, J. B. Shurin, and J. E. Smith. 2011.

Nutrient co‐limitation of primary producer communities. Ecology Letters 14:852–

862.

Harpole, W. S., and D. Tilman. 2007. Grassland species loss resulting from reduced niche

dimension. Nature 446:791–793.

Hays, C. G. 2007. Adaptive phenotypic differentiation across the intertidal gradient in the

alga Silvetia compressa. Ecology 88:149–157.

Helmuth, B., J. G. Kingsolver, and E. Carrington. 2005. Biophysics, physiological

ecology, and climate change: Does mechanism matter? Annual Review of

Physiology 67:177–201.

Hothorn, T., F. Bretz, and P. Westfall. 2008. Simultaneous Inference in General

Parametric Models. Biometrical Journal 50:346–363.

Huston, M., and D. DeAngelis. 1994. Competition and Coexistence - the Effects of

Resource Transport and Supply Rates. American Naturalist 144:954–977.

Huyer, A. 1983. Coastal Upwelling in the California Current System. Progress in

Oceanography 12:259–284.

Iles, A. C., T. C. Gouhier, B. A. Menge, J. S. Stewart, A. J. Haupt, and M. C. Lynch.

2012. Climate‐driven trends and ecological implications of event‐scale upwelling

in the California Current System. Global Change Biology 18:783–796.

Jones, M. N. 1984. Nitrate reduction by shaking with cadmium: Alternative to cadmium

columns. Water Research 18:643–646.

Kavanaugh, M. T., K. J. Nielsen, F. T. Chan, B. A. Menge, R. M. Letelier, and L. M.

Goodrich. 2009. Experimental assessment of the effects of shade on an intertidal

kelp: Do phytoplankton blooms inhibit growth of open-coast macroalgae?

Limnology and Oceanography 54:276–288.

Keister, J. E., W. T. Peterson, and S. D. Pierce. 2009. Zooplankton distribution and cross-

shelf transfer of carbon in an area of complex mesoscale circulation in the

northern California Current. Deep-Sea Research Part I-Oceanographic Research

Papers 56:212–231.

Kirincich, A., J. Barth, B. Grantham, B. A. Menge, and J. Lubchenco. 2005. Wind-driven

inner-shelf circulation off central Oregon during summer. Journal of Geophysical

Research-Oceans 110.

Kirk, J. T. O. 1994. Light and Photosynthesis in Aquatic Ecosystems. Second edition.

Cambridge University Press, Cambridge.

Koenker, R. 2013. quantreg: Quantile Regression.

Larkum, A. W. D., R. J. Orth, and C. M. Duarte. 2006. Seagrasses: biology, ecology, and

conservation. Springer, Dordrecht, the Netherlands.

Page 173: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

155

LeBauer, D. S., and K. K. Treseder. 2008. Nitrogen limitation of net primary productivity

in terrestrial ecosystems is globally distributed. Ecology 89:371–379.

Legaard, K. R., and A. C. Thomas. 2006. Spatial patterns in seasonal and interannual

variability of chlorophyll and sea surface temperature in the California Current.

Journal of Geophysical Research: Oceans 111.

Lewis, J. R. 1964. The ecology of rocky shores. The English Universities Press, Limited,

London, England.

Liebig, J. F. von. 1842. Animal Chemistry: Or, Organic Chemistry in Its Application to

Physiology and Pathology. (W. Gregory, Ed.). John Owen.

Lindeberg, M. R., S. C. Lindstrom, and Alaska Sea Grant College Program. 2010. Field

guide to seaweeds of Alaska. University of Alaska Fairbanks, Alaska Sea Grant

College Program, Fairbanks, Alaska.

Lobban, C. S., and P. J. Harrison. 1994. Seaweed ecology and physiology. Cambridge

University Press, Cambridge [England]; New York, NY, USA.

Di Lorenzo, E., N. Schneider, K. M. Cobb, P. J. S. Franks, K. Chhak, A. J. Miller, J. C.

McWilliams, S. J. Bograd, H. Arango, E. Curchitser, T. M. Powell, and P.

Rivière. 2008. North Pacific Gyre Oscillation links ocean climate and ecosystem

change. Geophysical Research Letters 35.

Lubchenco, J. 1980. Algal zonation in the New England rocky intertidal community: An

experimental analysis. Ecology 61:333–344.

MacNally, R. 1996. Hierarchical partitioning as an interpretative tool in multivariate

inference. Australian Journal of Ecology 21:224–228.

Madsen, T. V., and S. C. Maberly. 1990. A comparison of air and water as environments

for photosynthesis by the intertidal alga Fucus spiralis (Phaeophyta). Journal of

Phycology 26:24–30.

Major, K. M., and I. R. Davison. 1998. Influence of temperature and light on growth and

photosynthetic physiology of Fucus evanescens (Phaeophyta) embryos. European

Journal of Phycology 33:129–138.

Mann, K. H. 1973. Seaweeds - Their Productivity and Strategy for Growth. Science

182:975–981.

Mantua, N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis. 1997. A Pacific

Interdecadal Climate Oscillation with Impacts on Salmon Production. Bulletin of

the American Meteorological Society 78:1069–1079.

Marleau, J. N., F. Guichard, F. Mallard, and M. Loreau. 2010. Nutrient flows between

ecosystems can destabilize simple food chains. Journal of Theoretical Biology

266:162–174.

Mazerolle, M. J. 2013. AICcmodavg: Model selection and multimodel inference based on

(Q)AIC(c).

McClatchie, S., R. Charter, W. Watson, N. Lo, K. Hill, M. Manzano-Sarabia, R.

Goericke, C. Collins, E. Bjorkstedt, F. B. Schwing, S. J. Bograd, M. Kahru, B. G.

Mitchell, J. A. Koslow, S. Ralston, J. Field, W. T. Peterson, R. Emmett, J.

Gomez-Valdes, B. E. Lavaniegos, G. Caxiola-Castro, L. Rogers-Bennet, J.

Gottschalck, M. L. Heureux, Y. Xue, L. Munger, G. Campbell, K. Merkens, D.

Camacho, A. Havron, A. Douglas, and J. Hildebrand. 2009. The State of the

Page 174: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

156

California Current, Spring 2008-2009 Cold Conditions Drive Regional

Differences in Coastal Production. California Cooperative Oceanic Fisheries

Investigations Reports 50:43–68.

Menge, B. A. 1992. Community Regulation: Under What Conditions Are Bottom-Up

Factors Important on Rocky Shores? Ecology 73:755–765.

Menge, B. A. 2004. Bottom-Up/Top-Down Determination of Rocky Intertidal

Shorescape Dynamics. Pages 62–81 in G. A. Polis, M. E. Power, and G. R. Huxel,

editors. Food Webs at the Landscape Level. University of Chicago Press,

Chicago, IL.

Menge, B. A., G. W. Allison, C. A. Blanchette, T. M. Farrell, A. M. Olson, T. A. Turner,

and P. van Tamelen. 2005. Stasis or kinesis? Hidden dynamics of a rocky

intertidal macrophyte mosaic revealed by a spatially explicit approach. Journal of

Experimental Marine Biology and Ecology 314:3–39.

Menge, B. A., C. Blanchette, P. Raimondi, T. Freidenburg, S. Gaines, J. Lubchenco, D.

Lohse, G. Hudson, M. Foley, and J. Pamplin. 2004. Species interaction strength:

Testing model predictions along an upwelling gradient. Ecological Monographs

74:663–684.

Menge, B. A., F. Chan, and J. Lubchenco. 2008. Response of a rocky intertidal

ecosystem engineer and community dominant to climate change. Ecology Letters

11:151–162.

Menge, B. A., F. Chan, K. J. Nielsen, E. Di Lorenzo, and J. Lubchenco. 2009. Climatic

variation alters supply-side ecology: impact of climate patterns on phytoplankton

and mussel recruitment. Ecological Monographs 79:379–395.

Menge, B. A., B. A. Daley, J. Lubchenco, E. Sanford, E. Dahlhoff, P. M. Halpin, G.

Hudson, and J. L. Burnaford. 1999. Top-down and bottom-up regulation of New

Zealand rocky intertidal communities. Ecological Monographs 69:297–330.

Menge, B. A., B. A. Daley, P. A. Wheeler, E. Dahlhoff, E. Sanford, and P. T. Strub.

1997. Benthic–pelagic Links and Rocky Intertidal Communities: Bottom-up

Effects on Top-down Control? Proceedings of the National Academy of Sciences

94:14530–14535.

Menge, B. A., T. C. Gouhier, S. D. Hacker, F. Chan, and K. J. Nielsen. In Prep.

Oceanographic and ecological drivers of spatial structure in a rocky intertidal

meta-ecosystem.

Menge, B. A., J. Lubchenco, M. Bracken, F. Chan, M. Foley, T. Freidenburg, S. Gaines,

G. Hudson, C. Krenz, H. Leslie, D. Menge, R. Russell, and M. Webster. 2003.

Coastal oceanography sets the pace of rocky intertidal community dynamics.

Proceedings of the National Academy of Sciences 100:12229–12234.

Menge, B. A., and D. N. L. Menge. 2013. Dynamics of coastal meta-ecosystems: the

intermittent upwelling hypothesis and a test in rocky intertidal regions. Ecological

Monographs 83:283–310.

Menge, B. A., and J. P. Sutherland. 1987. Community regulation - Variation in

disturbance, competition, and predation in relation to environmental-stress and

recruitment. American Naturalist 130:730–757.

Page 175: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

157

Naldi, M., and P. Viaroli. 2002. Nitrate uptake and storage in the seaweed Ulva rigida C.

Agardh in relation to nitrate availability and thallus nitrate content in a eutrophic

coastal lagoon (Sacca di Goro, Po River Delta, Italy). Journal of Experimental

Marine Biology and Ecology 269:65–83.

Nielsen, K. J. 2003. Nutrient loading and consumers: Agents of change in open-coast

macrophyte assemblages. Proceedings of the National Academy of Sciences

100:7660 –7665.

Nielsen, K. J., and S. A. Navarrete. 2004. Mesoscale regulation comes from the

bottom‐up: intertidal interactions between consumers and upwelling. Ecology

Letters 7:31–41.

Paine, R. T. 1966. Food web complexity and species diversity. The American Naturalist

100:65–75.

Peterson, W. T., R. Emmett, R. Goericke, E. Venrick, A. Mantyla, S. J. Bograd, F. B.

Schwing, R. Hewitt, N. Lo, and W. Watson. 2006. The state of the California

Current, 2005-2006: warm in the north, cool in the south. California Cooperative

Oceanic Fisheries Investigations Reports 47:30.

Peterson, W. T., and F. B. Schwing. 2003. A new climate regime in Northeast Pacific

ecosystems. Geophysical Research Letters 30:1896.

Pfetzing, J., D. B. Stengel, M. M. Cuffe, A. V. Savage, and M. D. Guiry. 2000. Effects of

temperature and prolonged emersion on photosynthesis, carbohydrate content and

growth of the brown intertidal alga Pelvetia canaliculata. Botanica Marina

43:399–407.

Pfister, C. A., and K. L. Van Alstyne. 2003. An Experimental Assessment of the Effects

of Nutrient Enhancement on the Intertidal Kelp Hedophyllum sessile

(Laminariales, Phaeophyceae). Journal of Phycology 39:285–290.

Phillips, J., and C. Hurd. 2003. Nitrogen ecophysiology of intertidal seaweeds from New

Zealand: N uptake, storage and utilisation in relation to shore position and season.

Marine Ecology Progress Series 264:31–48.

Pinheiro, J., D. Bates, S. DebRoy, D. Sarkar, and R Development Core Team. 2013.

nmle: Linear and Nonlinear Mixed Effects Models.

Polis, G. A., W. B. Anderson, and R. D. Holt. 1997. Toward an Integration of Landscape

and Food Web Ecology: The Dynamics of Spatially Subsidized Food Webs.

Annual Review of Ecology and Systematics 28:289–316.

Polis, G. A., and S. D. Hurd. 1996. Linking Marine and Terrestrial Food Webs:

Allochthonous Input from the Ocean Supports High Secondary Productivity on

Small Islands and Coastal Land Communities. The American Naturalist 147:396–

423.

R Core Team. 2013. R: A Language and Environment for Statistical Computing. R

Foundation for Statistical Computing, Vienna, Austria.

Rahel, F. J., and W. A. Hubert. 1991. Fish assemblages and habitat gradients in a Rocky

Mountain–Great Plains stream: biotic zonation and additive patterns of

community change. Transactions of the American Fisheries Society 120:319–332.

Ramirez-Garcia, P., A. Lot, C. M. Duarte, J. Terrados, and N. S. R. Agawin. 1998.

Bathymetric distribution, biomass and growth dynamics of intertidal Phyllospadix

Page 176: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

158

scouleri and Phyllospadix torreyi in Baja California (Mexico). Marine Ecology

Progress Series 173:13–23.

Reiners, W. A., and G. E. Lang. 1979. Vegetational Patterns and Processes in the Balsam

Fir Zone, White Mountains New Hampshire. Ecology 60:403–417.

Robles, C. D., R. A. Desharnais, C. Garza, M. J. Donahue, and C. A. Martinez. 2009.

Complex equilibria in the maintenance of boundaries: experiments with mussel

beds. Ecology 90:985–995.

Robles, C., and R. Desharnais. 2002. History and current development of a paradigm of

predation in rocky intertidal communities. Ecology 83:1521–1536.

Roemmich, D., and J. A. McGowan. 1995. Climatic Warming and the Decline of

Zooplankton in the California Current. Science 267:1324–1326.

Sanford, E. 2002. The feeding, growth, and energetics of two rocky intertidal predators

(Pisaster ochraceus and Nucella canaliculata) under water temperatures

simulating episodic upwelling. Journal of Experimental Marine Biology and

Ecology 273:199–218.

Sanford, E., M. S. Roth, G. C. Johns, J. P. Wares, and G. N. Somero. 2003. Local

Selection and Latitudinal Variation in a Marine Predator-Prey Interaction. Science

300:1135–1137.

Sardans, J., A. Rivas-Ubach, and J. Peñuelas. 2012. The elemental stoichiometry of

aquatic and terrestrial ecosystems and its relationships with organismic lifestyle

and ecosystem structure and function: a review and perspectives.

Biogeochemistry 111:1–39.

Schwing, F. B., M. O’Farrell, J. M. Steger, and K. Baltz. 1996. Coastal upwelling

indices- West Coast of North America (1946-1995). NOAA National Marine

Fisheries Service.

Shanks, A. L., B. A. Grantham, and M. H. Carr. 2003. Propagule dispersal distance and

the size and spacing of marine reserves. Ecological Applications 13:159–169.

Sjøtun, K., S. Fredriksen, and J. Rueness. 1996. Seasonal growth and carbon and nitrogen

content in canopy and first-year plants of Laminaria hyperborea (Laminariales,

Phaeophyceae). Phycologia 35:1–8.

Smith, S. V. 1981. Marine Macrophytes as a Global Carbon Sink. Science 211:838–840.

Statzner, B., and B. Higler. 1986. Stream hydraulics as a major determinant of benthic

invertebrate zonation patterns. Freshwater Biology 16:127–139.

Stengel, D. B., and M. J. Dring. 1997. Morphology and in situ growth rates of plants of

Ascophyllum nodosum (Phaeophyta) from different shore levels and responses of

plants to vertical transplantation. European Journal of Phycology 32:193–202.

Stephenson, T. A., and A. Stephenson. 1949. The universal features of zonation between

tide-marks on rocky coasts. Journal of Ecology 37:289–305.

Sterner, R. W., and J. J. Elser. 2002. Ecological Stoichiometry: The Biology of Elements

from Molecules to the Biosphere. Princeton University Press, Princeton.

Sterner, R. W., J. J. Elser, E. J. Fee, S. J. Guildford, and T. H. Chrzanowski. 1997. The

Light: Nutrient Ratio in Lakes: The Balance of Energy and Materials Affects

Ecosystem Structure and Process. The American Naturalist 150:663–684.

Page 177: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

159

Sundareshwar, P. V., J. T. Morris, E. K. Koepfler, and B. Fornwalt. 2003. Phosphorus

Limitation of Coastal Ecosystem Processes. Science 299:563–565.

Tait, L. W., and D. R. Schiel. 2011. Dynamics of productivity in naturally structured

macroalgal assemblages: importance of canopy structure on light-use efficiency.

Marine Ecology Progress Series 421:97–107.

Van Tamelen, P. G., and M. S. Stekoll. 1997. The role of barnacles in the recruitment and

subsequent survival of the brown alga, Fucus gardneri (Silva). Journal of

Experimental Marine Biology and Ecology 208:227–238.

Van Tamelen, P. G., M. S. Stekoll, and L. Deysher. 1997. Recovery processes of the

brown alga Fucus gardneri following the “Exxon Valdez” oil spill: settlement and

recruitment. Marine Ecology Progress Series 160:265–277.

Tang, C. Q., and M. Ohsawa. 1997. Zonal transition of evergreen, deciduous, and

coniferous forests along the altitudinal gradient on a humid subtropical mountain,

Mt. Emei, Sichuan, China. Plant Ecology 133:63–78.

Thomas, T. E., P. J. Harrison, and D. H. Turpin. 1987. Adaptations of Gracilaria pacifica

(Rhodophyta) to nitrogen procurement at different intertidal locations. Marine

Biology 93:569–580.

Thornber, C., E. Jones, and J. Stachowicz. 2008. Differences in herbivore feeding

preferences across a vertical rocky intertidal gradient. Marine Ecology Progress

Series 363:51–62.

Tilman, D. 1985. The Resource-Ratio Hypothesis of Plant Succession. The American

Naturalist 125:827–852.

Tilman, G. D. 1984. Plant Dominance Along an Experimental Nutrient Gradient. Ecology

65:1445–1453.

Topinka, J. A., and J. V. Robbins. 1976. Effects of Nitrate and Ammonium Enrichment

on Growth and Nitrogen Physiology in Fucus spiralis. Limnology and

Oceanography 21:659–664.

Turner, T. 1985. Stability of Rocky Intertidal Surfgrass Beds: Persistence, Preempton,

and Recovery. Ecology 66:83.

US Department of Commerce, N. (n.d.). NOAA Earth System Research Laboratory.

http://www.esrl.noaa.gov/psd/enso/mei/.

Vitousek, P., and R. Howarth. 1991. Nitrogen Limitation on Land and in the Sea - How

Can It Occur. Biogeochemistry 13:87–115.

Wahl, M., V. Jormalainen, B. K. Eriksson, J. A. Coyer, M. Molis, H. Schubert, M.

Dethier, R. Karez, I. Kruse, M. Lenz, G. Pearson, S. Rohde, S. A. Wikström, and

J. L. Olsen. 2011. Chapter Two - Stress Ecology in Fucus: Abiotic, Biotic and

Genetic Interactions. Pages 37–105 Advances in Marine Biology. Academic

Press.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1999. Effects of resource

limitation on a detrital-based ecosystem. Ecological Monographs 69:409–442.

Wheeler, P. A., and B. R. Björnsäter. 1992. Seasonal fluctuations in tissue nitrogen,

phosphorus, and N: P for five macroalgae species common to the Pacific

Northwest coast. Journal of Phycology 28:1–6.

Page 178: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

160

Wieters, E. A., B. R. Broitman, and G. M. Branch. 2009. Benthic community structure

and spatiotemporal thermal regimes in two upwelling ecosystems: Comparisons

between South Africa and Chile. Limnology and Oceanography 54:1060–1072.

Williams, S. L., and M. N. Dethier. 2005. High and dry: variation in net photosynthesis

of the intertidal seaweed Fucus gardneri. Ecology 86:2373–2379.

Wootton, J. T. 1991. Direct and indirect effects of nutrients on intertidal community

structure: variable consequences of seabird guano. Journal of Experimental

Marine Biology and Ecology 151:139–153.

Worm, B., and A. R. O. Chapman. 1996. Interference competition among two intertidal

seaweeds: Chondrus crispus strongly affects survival of Fucus evanescens

recruits. Marine Ecology Progress Series 145:297–301.

Worm, B., H. Lotze, H. Hillebrand, and U. Sommer. 2002. Consumer versus resource

control of species diversity and ecosystem functioning. Nature 417:848–851.

Wright, J. T., S. L. Williams, and M. N. Dethier. 2004. No zone is always greener:

variation in the performance of Fucus gardneri embryos, juveniles and adults

across tidal zone and season. Marine Biology 145:1061–1073.

Zieman, J. C. 1974. Methods for the study of the growth and production of turtle grass,

Thalassia testudinum König. Aquaculture 4:139–143.

Zuccaro, A., C. L. Schoch, J. W. Spatafora, J. Kohlmeyer, S. Draeger, and J. I. Mitchell.

2008. Detection and Identification of Fungi Intimately Associated with the Brown

Seaweed Fucus Serratus. Applied and Environmental Microbiology 74:931–941.

Zuur, A. F., E. N. Ieno, N. J. Walker, A. A. Saveliev, and G. M. Smith. 2009. Mixed

effects models and extensions in ecology with R. Springer, New York; London.

Page 179: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

161

APPENDICES

Page 180: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

162

APPENDIX A – CHAPTER 2 SUPPLEMENTAL FIGURES & TABLES

Figure A2.1. Interannual variability in mean (± SE) macrophyte C:N for F. distichus

across regions from north to south (a: Northern Oregon, b: Southern Oregon). Capes

within each region are shown in filled and open circles, see Fig. 2.1 for cape codes.

Page 181: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

163

Figure A2.2. Interannual variability in mean (± SE) macrophyte C:N for M. splendens

across regions from north to south (a: Northern Oregon, b: Southern Oregon, c:

Northern California). Capes within each region are shown in filled and open circles;

see Fig. 2.1 for cape codes.

Page 182: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

164

Figure A2.3. Interannual variability in mean (± SE) macrophyte C:N for P. scouleri

across regions from north to south (a: Northern Oregon, b: Southern Oregon, c:

Northern California). Capes within each region are shown in filled and open circles;

see Fig. 2.1 for cape codes.

Page 183: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

165

Figure A2.4. Interannual variability in mean (± SE) macrophyte C:N for S. sessilis

across regions from north to south (a: Northern Oregon, b: Southern Oregon, c:

Northern California). Capes within each region are shown in filled and open circles;

see Fig. 2.1 for cape codes.

Page 184: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

166

APPENDIX B – CHAPTER 4 SUPPLEMENTAL FIGURES & TABLES

Figure B4.1. Occurrence of Fucus recruits (filled circles, #/0.25m2) and primary cover

of barnacles (open circles, % cover in 0.25m2) across vertical transects running from

+0.15m to +1.1m above MLLW.

Page 185: The Roles of Nutrient Dynamics, Oceanography, and Light in the Structure and Functioning of

167

Table B4.1. Results for drop-in-deviance tests comparing generalized linear models

including treatment nested within zone to models including only zone for field

experiment data. All results indicate that the reduced model is sufficient.

Response F-statistic P-value

Growth in length (cm) 0.683 0.686

New tip growth (#) 0.572 0.775

New receptacle formation (#) 0.268 0.963

Final receptacle proportion -2.937 0.891

% Carbon 0.451 0.834

% Nitrogen 0.776 0.600

Carbon:Nitrogen 0.530 0.777

*= Deviance, not F-statistic, was used