oceanography of the ross sea: ocean ...ocean.stanford.edu/people/dissertations/reddy...list of...

145
OCEANOGRAPHY OF THE ROSS SEA: OCEAN CIRCULATION, SEA ICE, AND PHYTOPLANKTON A DISSERTATION SUBMITTED TO THE DEPARTMENT OF ENVIRONMENTAL EARTH SYSTEM SCIENCE AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Tasha Elise Reddy June 2008

Upload: others

Post on 10-Jun-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

OCEANOGRAPHY OF THE ROSS SEA: OCEAN CIRCULATION, SEA ICE, AND PHYTOPLANKTON

A DISSERTATION SUBMITTED TO THE DEPARTMENT OF

ENVIRONMENTAL EARTH SYSTEM SCIENCE AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

Tasha Elise Reddy

June 2008

Page 2: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

UMI Number: 3313648

Copyright 2008 by

Reddy, Tasha Elise

All rights reserved.

INFORMATION TO USERS

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

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

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

alignment can adversely affect reproduction.

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

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

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

®

UMI UMI Microform 3313648

Copyright 2008 by ProQuest LLC.

All rights reserved. This microform edition is protected against

unauthorized copying under Title 17, United States Code.

ProQuest LLC 789 E. Eisenhower Parkway

PO Box 1346 Ann Arbor, Ml 48106-1346

Page 3: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Copyright © 2008 by Tasha Elise Reddy

All Rights Reserved

11

Page 4: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Kevin R. Arrigo, Principal Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Oliver B. Fringes:

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

)avid M. Holland

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for die degree of Doctor of Philosophy.

Adina Paytan I /

Approved for the Stanford University Committee on Graduate Studies.

@*z-f-. rf^ph-

in

Page 5: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

ABSTRACT

It is a crucial time to further our understanding of oceanographic systems, especially at

high latitudes, as these regions are expected to be the earliest and most susceptible sites of future

change. The Ross Sea, Antarctica, makes an ideal focus area, as its physical and biological systems

are well studied, but gaps in our knowledge still remain. To gain insight into the Ross Sea system,

I used a number of different methods, primarily targeting improved understanding of ocean

circulation in the region. These methods include analysis of ocean color satellite imagery, model

simulations of oceanic chlorofluorocarbons, and model simulations of sea ice involving the

formation and evolution of the Ross Sea polynya. The simulations are run with ocean circulation

models coupled to sea ice and glacier components with atmospheric forcing. I find that warm,

deep water comes from offshore onto the continental shelf as Modified Circumpolar Deep Water

(MCDW), retarding the growth of sea ice and also entering the Ross Ice Shelf (RIS) cavity where

it contributes to basal melting. The geostrophic nature of the flow on the shelf constrains the

movement of waters onshore and causes them to be tightly coupled to bathymetric features.

Offshore waters limit the extent of the phytoplankton bloom which is contrary to previous

suggestions that nutrient-rich MCDW mixes into the surface in the summer, enhancing the

bloom. There has been some controversy in the past as to the degree that different factors

contribute to the formation of the polynya in the Ross Sea. I find that winds play the dominant

role in its formation and there is a contribution from ocean heat. I additionally predict that basal

melting of the RIS is highest in the early spring and early autumn. Little is known about the

seasonal melting of the RIS, and this research helps create a clearer understanding of seasonally

changing mass-balance. The early spring melt can likely be attributed to warm MCDW entering

the RIS cavity year-round and then mixing into the surface waters during the winter. Finally, I

look at future change scenarios, completing 50-year coupled ocean, sea ice, atmosphere model

runs, studying sea ice concentrations and primary productivity under altered atmospheric forcing

of wind and air temperatures. In contrast to earlier findings that predict substantial changes, my

model runs estimate that in the next 50 years, sea ice concentrations and primary production in

the Ross Sea will change a relatively small amount. I find that under both temperature and wind

forcing scenarios, sea ice decreases in the summer and the winter, but does not disappear. This

does not eliminate the potential of a tipping point to occur after the 50-year run which could

result in a dramatically non-linear response of sea ice to altered atmospheric forcing. I find that

primary production changes in magnitude and timing in these scenarios, but does not change

significantly on average. This does not rule out that there may be plankton community shifts due

to changing physical conditions.

IV

Page 6: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

ACKNOWLEDGMENTS

I would like to thank my advisor, Kevin Arrigo, for many years of guidance and

inspiration while completing my degree. As a great researcher, teacher, and mentor, he has

helped me to develop into the scientist I am today. Additionally, Kevin has demonstrated how to

balance the successful management of a large research group with making time to spend outside

of the office with his family.

I would like to thank David Holland. It has been a pleasure to collaborate with him over

the years and I look forward to working with him in years to come. David has an admirable way

of making science fun and his enthusiasm for his own work, and science in general, is vast.

My two other committee members, Oliver Fringer and Adina Paytan, have also provided

inspiration to me during the completion of my graduate studies. Oliver, as a junior faculty

member, has been a role model to me as a young scientist. Adina, as the professor of many of my

courses, exudes scientific knowledge and wisdom. She is also an amazing example of how to

generously take time to reach out to the community and offer what she has as a scientist in order

to advance the development of the next generation.

And finally I thank my committee chair, Terry Root. I was very fortunate to meet Terry

during a Stanford short-course on Global Warming. Simply put, this was the best course that I've

ever participated in in my life. In one brief week, Terry provided me with lasting mentorship and

was an incredible role model as a woman in a senior faculty of science position.

For helpful discussions, comments, and advice in preparation of the manuscripts

presented within this thesis, I would like to again thank Kevin Arrigo, David Holland, and Oliver

Fringer, and additionally thank Derek Fong, Rochelle Labiosa, Steven Monismith, Stan Jacobs,

and William Smethie. I would also like to acknowledge the multiple anonymous reviewers and

editors for their valuable suggestions during the submissions of the manuscripts presented here.

Page 7: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

I thank the graduate students of the ocean group at Stanford. Over the years they have

provided much support and camaraderie. In particular, I thank the members of the Arrigo

research group: Rochelle Labiosa, Lindsey Kropuenske, Gert van Dijken, Sudeshna Pabi, Mine

Berg, Anne-Carlijn Alderkamp, Ben Saenz, Matt Mills, Viola Toniolo, Molly Palmer, and Stuart

Borrett. I have very much enjoyed learning from and spending time with all of them during my

time at Stanford.

Finally I thank my family and friends, especially Lara, Erin, Matt, Brennan, Carol, Alexis,

Bohdan, Mike, Herb, Ayyana, Will, Kelly, Jason, Cat, Ryan, Courtney, Kamini, CJ, Steve, Martine,

Brad, Lindsey, Chi Soo, Eugene, Darcy, Rick, Nick, Brie, James, Malita, Jim, Phoebe, Anna, Thijs,

Matt, Eve, Jenny, Mine, Rochelle, Janine, Scott, Laura, Mike, Farshad, Tricia, Matt, Dave and

Dave, Sang-Ho, Charlie, Seth, Wendy, Sue, Kerry, Susan, Serkan, Laura, Jess, Peter, Stephen,

Philip, and William. Thank you Marge and grandma, and thank you Biggie for your guidance and

inspiration. I wish you could be here to witness my accomplishment and share in the celebration.

In particular, I thank my mother who has been a constant pillar of support for me, unwavering

over the years. I will forever be in debt to her for her love and patience. Most of all, I thank

Adam. Without him, I would not be who I am or where I am today. My life has presented me

with great challenges over the past few years, and Adam has been there for me at every turn along

the way. With all my heart, I thank him for being my best friend.

This work was supported by NASA Earth System Science Fellowship grant NGT5-

30489, Stanford University McGee grants, and Shell Foundation awards. Additional support

came from NASA, DOE, and NSF grants to Kevin Arrigo and David Holland.

V I

Page 8: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

TABLE OF CONTENTS

ABSTRACT iv

ACKNOWLEDGMENTS v

TABLE OF CONTENTS vii

LIST OF TABLES ix

LIST OF FIGURES x

CHAPTER 1: Introduction 1

CHAPTER 2: Constraints on the Extent of the Ross Sea Phytoplankton Bloom 6

Abstract 6

2.1 Introduction 7 2.1.1 Remote Sensing 11 2.1.2 Modeling 12

2.3 Results 13 2.4 Patterns in Chi and Mechanisms for their Formation 16 2.5 Advection via Ekman Layer, Tidal, and Geostrophic Flow 20 2.6 Conclusions 22

CHAPTER 3: The Role of Thermal and Mechanical Processes in the Formation of the Ross Sea Summer Polynya 24

Abstract 24 3.1 Introduction 25 3.2 Methods 28

3.2.1 Model Description 28 3.2.2 Control Run and Parameter Sensitivity 32 3.2.3 Model Configuration Sensitivity Runs 34

3.3 Results 37 3.3.1 Control Run 37 3.3.2 Sensitivity Studies 43

3.3.2.1 Sensitivity to Parameters Values 43 3.3.2.2 Sensitivity to Model Structure 43 3.3.2.3 Sensitivity to Ocean Heat 44 3.3.2.4 Sensitivity to Ice Advection 48 3.3.2.5 Sensitivity to Winds 49

3.4 Discussion 51 3.5 Conclusions 58

vn

Page 9: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 4: Ross Ice Shelf Cavity Circulation, Residence Time, and Melting: Results from a Model of Oceanic Chlorofiuorocarbons 59

Abstract 59 4.1 Introduction 60 4.2 Modeling Methods 63

4.2.1 Ocean Component 63 4.2.2 Sea-Ice and Ice Shelf Components 64 4.2.3 Atmosphere and CFC Gas Exchange 65 4.2.4 Model Grid and Runtime 67 4.2.5 Residence Time 68

4.3 Results and Discussion 68 4.3.1 Modeled CFC Distributions 68 4.3.2 Modeled Circulation, Salinity, and Temperature 72 4.3.3 Residence Time Estimates 76 4.3.4 Low-CFC Waters 76 4.3.5 Basal Melt Rates 79

4.4 Conclusions 81

CHAPTER 5: Ross Sea Primary Production and Sea Ice in Future Climatic Change Scenarios 83

Abstract 83 5.1 Introduction 84 5.2 Methods 86

5.2.1 The Physical Model 86 5.2.2 The Biological Model 88 5.2.3 Climate Change Scenarios 90

5.3 Results 93 5.3.1 Sea Ice 93 5.3.2 Ocean Temperatures 98 5.3.3 Primary Productivity 101 5.3.4 Sensitivity to the Magnitude of Changes 104

5.4 Discussion 107 5.4.1 Southern Ocean Sea Ice 107 5.4.2 Arctic and Antarctic Differences 109 5.4.3 Response of Biological Systems to Changes in Physical Systems I l l

5.5 Conclusions 114

CHAPTER 6: Conclusions 116

REFERENCES 120

viii

Page 10: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

LIST OF TABLES

Table 3.1 Uncertain model parameters and coefficients used in sensitivity analyses 33

Table 3.2 Model configurations for sensitivity runs 34

Table 3.3 Model sensitivity analyses 36

Table 4.1 Chlorofluorocarbon descriptions 60

Table 4.2 Constants for the temperature dependent Schmidt number [Zheng 1998 (CFC-11 and CFC-12); Haine and Richards, 1995 (CFC-113)] and gas solubility [in units of pmol kg1 patnY1, Warner and Weiss 1985 (CFC-11 and CFC-12); Bu and Warner 1995 (CFC-113)] 67

Table 5.1 Comparison of peak daily, mean daily, and annual primary productio 101

IX

Page 11: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

LIST OF FIGURES

Figure 2.1 (a) Map of the Antarctic coastline and 1000 m bathymetric contour with study region outlined in blue. A schematic of the Ross Sea Gyre is shown in green, (b) Bathymetry map for the study area also including Ross Sea Gyre circulation schematic (green). Purple line marks the boundaries of images displayed in Figure 2 7

Figure 2. 2 Satellite imagery of chlorophyll from December (a) 1978, (b)1998, (c) 1999, and (d) 2003, and (e) bathymetry. Areas of low chlorophyll associated with the bloom are indicated by white lines (1-3). Areas of high-chlorophyll waters between the jets of low-chlorophyll waters are labeled A through C. The labels are included on all images regardless of whether the feature is present or not during that year. Grey areas indicate sea ice and black regions represent areas of missing satellite ocean color data, most often due to cloud cover. Note that in 2003 (d), the western boundary of the bloom is located westward of the usual location (see text) 8

Figure 2.3 Instantaneous surface velocity streamlines overlain on bathymetry 13

Figure 2.4 Comparison of December chlorophyll distributions from (a) model results and (b) SeaWiFS imagery from 1998 14

Figure 2.5 Model results showing surface (a) salinity and (b) temperature and vertical profiles along a transect of 74.8°S of (c) salinity and (d) temperature 15

Figure 2.6 Modeled results for surface waters in December of (a) chlorophyll and (b) iron 18

Figure 2.7 Instantaneous surface velocity streamlines overlain on modeled surface density anomaly. 22

Figure 3.1 Map of the Antarctic coastline and the 1000 m bathymetric contour. The Ross Sea model domain perimeter is shown (heavy outline) as well as the perimeter for the model domain which excludes the Ross Ice Shelf (dashed line) 30

Figure 3. 2 Bathymetry (m) of the Ross Sea model domain which includes the Ross Ice Shelf cavity, the continental shelf and the abyssal ocean. Also shown are the three regions for which statistics were extracted 31

Figure 3.3 Sea ice concentrations from (a) satellite climatology (1987-1993) and (b) year 10 of the model control run 38

Figure 3.4 Sea-ice thicknesses from (a) June 1995 data (b) June 1998 data (c) June 1st

model control run (d) January 1999 data (e) January 2000 data (£) Feb 1st

x

Page 12: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

model control run (sea-ice thickness data compiled by DeUberty and Geiger [2005]) 39

Figure 3. 5 Comparison of ice and snow thicknesses (a) from May/June 1995 field observations along transects between 160°W and 180° [Jeffries andAdolphs, 1997] and (b) from June of the 10th year of the model control run along a 180° transect (c) November 1998 field observations along a transect of approximately 175°E [Arrigo eta/., 2003] and (d) from November of the 10th

year of the model control run along a 175°E transect 41

Figure 3.6 Austral summer sea-ice concentrations of the 10th year of the model run from (a) model control run (in which the model domain excludes the cavity under the RIS) (b) no heat entrainment June - March (c) no heat entrainment December - March (d) no heat entrainment for the entire 10 year spin-up period (e) ice not moving June - March (f) no heat entrainment and ice not moving June - March (g) open water albedo feedback effect from June - March (h) no wind from June - March (i) no wind from December - March Q) no snow from June — March (k) no snow from December - March 45

Figure 3.7 Percent of time in November and December during which ice concentrations were greater than 10% (satellite data from 1997 - 2001). The position of the 1000m isobath is shown by the black line. [Arrigo and Van Dijken, 2004] 54

Figure 3.8 Modeled (a) vertical temperature profiles at 175°E and (b) sea surface temperatures over the model domain 56

Figure 4.1 Atmospheric CFC concentrations in the Southern Hemisphere of CFC-12, CFC-11, and CFC-113 from 1940 to 2000. [Walker et al. 2000] 61

Figure 4. 2 (a) Map of the Antarctic coastline and the 1000 m bathymetric contour. The Ross Sea model domain perimeter is shown, (b) Sampling locations of CFC field data along transect following the RIS edge [Smethie and Jacobs 2005] during Jan-Feb 1984, Feb 1994, and Feb-Mar 2000 62

Figure 4.3 CFC-12 concentrations [pmol kg"1] from (a) field studies along a transect following the ice shelf edge Jan-Feb 1984; Feb 1994; Feb-Mar 2000 [Smethie and Jacobs 2005] (b) model estimates along a transect of 77.3°S for equivalent model dates (c) model estimates for the ocean at 450 m depth surrounding the mouth of the cavity (d) model estimates for the ocean at 450 m depth (for the entire model domain) (e) model estimates for the ocean mixed layer (for the entire model domain) 70

Figure 4.4 CFC-12 [pmol kg"1] concentration model estimates in depth and time at a station 180°, 77.3°S for 1980 through 2000 71

XI

Page 13: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Figure 4. 5 Modeled temperature and velocity (averaged over the final year of the model run) for (a) surface layer and (b) at 450m depth 72

Figure 4.6 Salinity from (a) field studies (along a transect following the ice shelf edge Jan-Feb 1984; Feb 1994; Feb-Mar 2000) and (b) model estimates (along a transect of 77.3°S for equivalent model dates) 74

Figure 4. 7 Temperature from (a) field studies (along a transect following the ice shelf edge Jan-Feb 1984; Feb 1994; Feb-Mar 2000) and (b) model estimates (along a transect of 77.3°S for equivalent model dates) (c) model estimates surrounding the mouth of the cavity 75

Figure 4.8 Distribution of RIS cavity "dye" in the mixed layer at the half-life of the dye in the cavity [%] 77

Figure 4.9 Modeled estimates of RIS basal melt rate [cm/a] (a) March (b) June (c) September (d) December 79

Figure 5.1 Increases in (a) temperatures and (b) winds over the 50 years of the model runs 92

Figure 5.2 Sea ice concentration (averaged over the model domain) for the 50 years of the model runs for (a) control run (b) intermediate temperature run (c) intermediate wind run and (d) temperature+wind run 94

Figure 5.3.1 Sea Ice Concentrations (Tight grey is open water; dark grey is the continent; light blue is the RIS) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run 95

Figure 5.4 Ocean mixed layer temperatures (dark grey is the continent; light blue is the RIS) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run 99

Figure 5.5 Primary Production (dark grey is the continent) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run 102

Figure 5.6 Primary production time series 103

Figure 5.7 Model results from the low, intermediate, and high model scenarios for temperature and wind of (a) January sea ice concentrations (b) February ocean mixed layer temperatures and (c) December (peak) primary production (light grey is open water; dark grey is the continent; light blue is the RIS) 105

xn

Page 14: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 1 Introduction

Even the seemingly isolated waters of the Southern Ocean play a key role in the

carbon (C) cycle and affect climate on a global scale. The ocean is an important global C

reservoir that exchanges with the atmosphere at timescales on the order of one thousand

years [Sarmiento and I^eQuere, 1996], and the Southern Ocean accounts for a significant

amount of global carbon dioxide (CO^ sequestrationfGz/^/ra and Duffy, 2000].

Anthropogenic climate change affects the ocean-ice-atmospheric system, particularly at

high latitudes, and it is far from certain how the system will respond to global warming

and what feedbacks may develop in the future [Sarmiento et al, 1998]. One likely feedback

involves the alteration of C export to the deep ocean. Understanding present

mechanisms that are at work at regional scales as part of the ocean-ice-atmosphere system

is of particular importance if we hope to understand what factors may be important at

global scales in the future, as the Earth responds to warming.

In general phytoplankton chlorophyll concentrations are higher over the

continental shelves of the Southern Ocean than over the open ocean. The Ross Sea

polynya is the most productive region of the Southern Ocean (1.46 gC m 2 d"1 in January

and 151 gC m2 yr4 annual [Arrigo and Van Dijken, 2003]), an area that for the most part

has low productivity (4414 Tg C yr ! for all water south of 50°S, [Arrigo et al, 1998])

despite high levels of macronutrients (high nutrient low chlorophyll, or HNLC region).

Due to its high productivity and intense phytoplankton blooms, the Ross Sea polynya

supports a substantial food web base for higher trophic levels. Thus understanding the

1

Page 15: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Ross Sea biological oceanographic system has widespread implications. There is a water

mass found at depth around Antarctica call Circumpolar Deep Water (CDW). It comes

onto the continental shelves (sometimes as Modified CDW (MCDW) or Upper CDW

(UCDW)) where it impacts biological and physical systems. Along the western Antarctic

Peninsula, nutrient-rich UCDW is known to supply nutrients to the surface waters and

enhance the phytoplankton bloom \Pre%elin et al, 2000]. The extent of phytoplankton

blooms, aside from being limited by nutrient availability and light availability, is affected

by ocean circulation as well as the presence of sea ice.

Sea ice limits the amount of light that enters the waters below, inhibiting a

phytoplankton bloom from forming when sea ice is present. In addition to affecting light

availability for phytoplankton, sea ice plays an important role in the physical

oceanographic system as a barrier to air-sea exchange of heat and gases, deep water

formation and deep mixing. Sea ice can be affected, among other things, by winds, ocean

heat, ocean currents, tides, and air temperatures. Aside from being the most productive

region of the Southern Ocean, the Ross Sea polynya is also the largest in the world \Arrigo

and Van Dijken, 2003] so it is important to understand how different factors affect its

formation.

Understanding the mass-balance of Antarctic ice shelves is in particularly key as

we continue in a world of unprecedented climatic change. While the ice shelves are

floating and do not contribute directly to sea level rise, they hold back continental ice

from flowing into the ocean. This ice can have a substantial impact on sea level. When

studying this mass balance, it is critical to quantify basal melting - melting along the base

of the ice shelf. Just as the Ross Sea polynya is the largest in the world, the cavity beneath

2

Page 16: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

the Ross Ice Shelf (SIS) is the largest ice shelf cavity in the world. Relatively little is know

about the base of the RIS and die cavity below due to die difficulty in making direct

measurements in such a remote location, but understanding this system is becoming

increasingly important.

Atmospheric temperatures have been [Brohan et al, 2006] and will be [IPCC 2007]

increasing, resulting in widespread impacts on almost all earth systems. Changes are

expected to occur earliest and more intensely at the poles [Flato et al, 2000; Shindell et al,

1999; Stouffer et al, 1989]. The Arctic has already undergone significant change with

dramatic declines in sea ice in the last decade. Additionally, high latitude precipitation is

expected to increase [Groisman et al, 2005] and winds are expected to shift in location and

change in intensity \Russell et al, 2006]. In order to predict and prepare for changes that

may occur in the future, forecast models are necessary to simulate future conditions.

For this study, two different regional models of the Ross Sea were used. The

Coupled Ice Atmosphere Ocean (CIAO) model simulates the Ross Sea ecosystem \Arrigo

et al, 2003; Worthen et al, 2003]. The physics in CIAO are based the sigma-coordinate

Princeton Ocean Model (POM) \£>lumberg, 1987]. It was modified to include a primary

production algorithm and biological variables such as nutrients (NOs, Si, P0 4 , and Fe),

zooplankton, detritus, and phytoplankton taxonomic composition and abundance [Arrigo

2003a]. The CIAO model takes into account the different C to P nutrient removal ratios

of the two dominant phytoplankton taxa in the region, Phaeocystis antarctica and diatoms.

The second model used is the Polar Ocean Land Atmosphere Regional modeling system

(POLAIR) [Holland and Jenkins, 2001; Holland et al, 2003; Reddy et al, 2007]. The ocean

component is an isopycnic layer model, based on the Miami Isopycnic Coordinate Model

3

Page 17: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

(MICOM) [Bleck and Smith, 1990], the sea ice component uses a cavitating fluid rheology,

and the simple atmospheric component uses daily forcings. The simulations incorporate

the cavity beneath the RIS. It has been modified as part of this dissertation to include a

biological component consisting of a phytoplankton growth model that includes nutrient

(Fe), phytoplankton (P. antarcticd), and detritus [Reddy andArrigo, in prep].

The focus of the research presented here is to understand the physical

oceanographic system and its impact on ecosystems in the Ross Sea, Antarctica. This

includes the circulation of ocean waters and the presence of sea ice, and thus, the

availability of sunlight for photosynthesis. In Chapter 2,1 document the existence of

large (100 km) features of high and low biomass regions in the phytoplankton bloom.

Chapter 2 has been published in tht Journal of Geophysical Research [Reddy andArrigo, 2006].

Chapter 3 documents processes that contribute to the formation of the Ross Sea polynya.

Specifically, we model the annual evolution of the polynya and analyze the response to

shutting down processes, one at a time, such as winds, tides, ocean currents, and ocean

heat. Chapter 3 has been published in the Journal of Geophysical Research [Reddy, Arrigo, and

Holland, 2007]. In Chapter 4, we us a model of oceanic chlorofluorocarbons to validate

estimates of circulation on the Ross Sea continental shelf and within the RIS cavity. We

then use this model to estimate the residence time of water within the RIS cavity as well

as the basal melt rates along the base of the RIS. Additionally, we use the modeled CFCs

to study the circulation of MCDW on the continental shelf. This chapter, Chapter 4, is

being submitted for publication to Limnology and Oceanography [Reddy, Holland, andArrigo, to

be submitted]. I also examine the potential changes to sea ice and primary production in

the future due to altered atmospheric forcing of wind and temperature. For the study in

4

Page 18: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Chapter 5,1 modeled 50 year scenarios of the future with wind speeds 10%, 50% and

100% greater than the control and also with air temperature change increasing to between

1°C and 4.5°C in 50 years. The impact of these changes is analyzed for seasonal sea ice

concentrations and spring and summer primary productivity. Chapter 5 is in preparation

for submission to Journal of Geophysical Research [Reddy andA.rrigo, in prep].

5

Page 19: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 2 Constraints on the Extent of the Ross Sea Phytoplankton Bloom

Abstract

An intense phytoplankton bloom forms over the continental shelf in the Ross Sea,

Antarctica every spring as the sea ice retreats during the expansion of the Ross Sea polynya.

The extent of this bloom rarely covers the full area of the shelf for reasons that are not well

understood. One clue may be found in satellite imagery spanning the last twenty-five years

that shows frequently recurring, previously undescribed, large scale plumes of high-

chlorophyll water and jets of low-chlorophyll water demarking the northern boundary of the

spring phytoplankton bloom. The plumes, with horizontal length scales of 100 — 200 km,

appear at the same location each year. The spatial consistency of this pattern of jets and

plumes suggests regulation by forcing that does not vary from year to year, such as the

interaction between bathymetry and prevailing ocean currents. Here we present a modeling

study indicating that this pattern in the spring phytoplankton bloom is caused by intrusions

of offshore surface waters onto the continental shelf, guided by relatively deep polar shelf

bathymetry. Our study suggests that currents from the Ross Sea gyre follow along the shelf

break and prevent the phytoplankton bloom from extending further towards the edge of the

shelf, even when the waters there are ice free, perhaps because the offshore waters have low

iron concentrations. These results illustrate how phytoplankton distributions can be

constrained by the interaction of deep bathymetry and surface geostrophic currents along the

highly productive Antarctic continental shelf.

6

Page 20: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

2.1 Introduction

Phytoplankton chlorophyll a (Chi) concentrations in the Southern Ocean are

generally higher on continental shelves, likely due to the greater abundance of iron in

shallowwaters [Comiso eta/., 1993]. The Ross Sea continental shelf (Figure 2.1) is no

exception, containing some of the most biologically productive waters in Antarctica

[Arrigo and Van Dijken, 2003b]. Phytoplankton biomass often exceeds 10 mg Chi m~3

[Smith and Gordon, 1997] during intense spring phytoplankton blooms that form when

light availability in surface waters increases due to the expansion of the Ross Sea polynya

(an area of open water surrounded by sea ice).

While the extent of the sea ice is one factor constraining the ultimate size of the

phytoplankton bloom, even when the continental shelf is virtually ice free, the bloom

rarely extends to the outer reaches of the shelf. Furthermore, the extent of the bloom

can vary dramatically from year to year \Arrigo et a/., 1998; Arrigo and Van Dijken, 2004].

Figure 2.1 (a) Map of the Antarctic coastline and 1000 m bathymetric contour with study region outlined in blue. A schematic of the Ross Sea Gyre is shown in green, (b) Bathymetry map for the study area also including Ross Sea Gyre circulation schematic (green). Purple line marks the boundaries of images displayed in Figure 2.2.

7

Page 21: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

170°E 175°E 180°E

-1000 -800 -600 -400 -200 0

Figure 2.2 Satellite imagery of chlorophyll from December (a) 1978, (b)1998, (c) 1999, and (d) 2003, and (e) bathymetry. Areas of low chlorophyll associated with the bloom are indicated by white lines (1-3). Areas of high-chlorophyll waters between the jets of low-chlorophyll waters are labeled A through C. The labels are included on all images regardless of whether the feature is present or not during that year. Grey areas indicate sea ice and black regions represent areas of missing satellite ocean color data, most often due to cloud cover. Note that in 2003 (d), the western boundary of fhe bloom is located westward of the usual location (see text).

Clues to what is limiting the spatial

extent of the phytoplankton bloom

may be found in ocean color satellite

imagery of the region collected between

1978 and 2003 \Arrigo and Van Dijken,

2004; Arrigo eta/., 1998] which show

large plumes of high-Chl water (Figure

2.2, areas A through C) separated by

regions of low-Chl water (Figure 2.2:

lines 1 through 3) that develop during

the peak of the spring bloom. The area

of low-Chl along the northwestern

boundary of the bloom (line 1) is not

spatially fixed from year to year and

may be associated with the position of

8

Page 22: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

the retreating ice edge, explaining its absence in 2003 (Figure 2.2d). The other two jets of

low-Chl (lines 2 and 3), however, are quite consistent from year to year. They flank both

sides of the Pennell Bank, suggesting a correlation between low surface Chi and deep

bathymetry. During years when bloom extent is larger than average, the eastern-most jet

of low Chi (line 3) does not appear. However, virtually every year that the southwestern

Ross Sea was ice free during spring (large blooms did not develop during some years

because of heavy sea ice due to icebergs or unusual climate conditions \Arrigo and Van

Dijken, 2003a; Arrigo and Van Dijken, 2004]), the center jet of low Chi (line 2) bordered by

two plumes of high Chi is spatially fixed and strikingly apparent in satellite imagery.

The formation of these mesoscale features is likely driven by surface circulation in

this region of the Southern Ocean. Circulation in the southwestern Ross Sea is

dominated by the swift surface-intensified slope current \Assmann eta/., 2003], the

southern part of the cyclonic (clockwise) Ross Sea Gyre that runs northwest along the

shelf break (Figure 2.1). Some of this water flows onto the Ross Sea continental shelf,

which averages approximately 600 m in depth but which has three shoals 300 m deep and

a steep shelf break (Figure 2.1). Previous modeling studies have found that the

circulation on the shelf is strongly affected by this irregular bottom topography [e.g.

Dinniman eta/., 2003].

There are a variety of ways in which relatively deep bathymetry can alter surface

circulation \Heath, 1981; Holland, 2001; Semtner and' Chervin, 1992] and affect

phytoplankton distributions. The most common mechanism is the topographically

enhanced entrainment of nutrients into surface waters and subsequent stimulation of

phytoplankton growth. This process has been observed in many locations, including the

9

Page 23: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

western Antarctic Peninsula \Pre%elin et al, 2000], the Subtropical Convergence at

Chatham Rise [McClatchle et al, 1997], and upwelling filaments off of the coast of

California \Flament et al, 1985] and southwest Africa [Shillington et al, 1990]. However, the

upwelling of nutrients does not likely affect phytoplankton distributions in the Ross Sea,

as springtime nutrients (including micronutrients such as iron) are generally high on the

continental shelf due to convective overturning during winter. Another way that

phytoplankton distributions are affected by deep bathymetry is through lateral advection

of surface waters that are bathymetrically constrained to create heterogeneous patterns in

surface phytoplankton distributions. For example, Moore et al. [1993] observed a non-

upwelling, bathymetrically constrained, offshore-directed plume of high Chi near the

coast of New Zealand. However, there are few other examples of such features

occurring along non-upwelling coasts. Finally, the depth of the well-mixed surface layer

can be altered as it flows over irregular bathymetry [Stewart, 2004]. These changes in

mixed layer depth (MLD) can alter light availability, and consequently, phytoplankton

growth rates and abundance.

The purpose of the present study was to investigate the physical processes that

control the size and shape of the phytoplankton bloom in the southwestern Ross Sea. In

particular, we were interested in whether the surface expression of the bloom is

bathymetrically constrained, and if so, by what mechanism. This was investigated using a

3-dimensional ocean ecosystem of the Ross Sea, specifically, the Coupled Ice,

Atmosphere, Ocean (CIAO) model \Arrigo et al, 2003; Worthen andArrigo, 2003]. This

model simulates phytoplankton dynamics as a function of seasonal changes in

temperature, light, and nutrient distributions, which are in turn controlled by patterns of

10

Page 24: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

advection and vertical mixing. By determining the conditions under which CIAO could

accurately simulate the major distributional patterns of the phytoplankton bloom, we

were able to assess whether the observed mesoscale features form regularly due to

bathymetric control of either advective flow or mixed-layer depth. While the features

themselves are not likely to have a significant effect on carbon uptake or determine upper

trophic level foraging patterns, identifying the mechanisms of their formation will help us

understand how mesoscale physical oceanographic processes constrain phytoplankton

distributions on this highly productive continental shelf.

2.2 Methods

2.1.1 Remote Sensing

Spatial patterns in Chi were derived from Sea-viewing Wide Field-of-view Sensor

(SeaWiFS 1997-2003) and Coastal Zone Color Scanner (CZCS 1978-1981) satellite

imagery obtained from the Goddard Earth Sciences Data and Information Services

Center (DAAC). We processed SeaWiFS data using the NASA SeaDAS image

processing software and the OC4v4 algorithm. CZCS pigment concentrations were

calculated using the standard algorithm of Gordon etal. [1983]. Despite heavy cloud cover

on average, clear images were obtained during ice free years with which to characterize

patterns of Chi and their relationship to bathymetry so that averaging over multiple days

was not necessary. Sea ice distributions were determined from daily Special Sensor

Microwave Imager (SSM/I) data available at the EOS Distributed Active Archive Center

(DAAC) at the National Snow and Ice Data Center, University of Colorado, Boulder,

CO.

11

Page 25: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

2.1.2 Modeling

To investigate possible interactions between surface phytoplankton blooms and

bathymetry, we increased the horizontal resolution of CIAO, a complete description of

which can be found in Arrigo eta/. [2003] and Worthen andArrigo [2003], from 25 km to 8

km in the southwestern Ross Sea. The physics of CIAO are based on the Princeton

Ocean Model (POM), a primitive equation, sigma-coordinate ocean circulation model

[B/umberg and Me/lor, 1987] with vertical mixing calculated using the turbulence closure

scheme of Mel/or and Yamada [1982]. CIAO specifies daily sea ice concentration from a

20-year climatology derived from passive microwave satellite data to calculate surface heat

and salt fluxes.

The biological component of the model includes the two dominant

phytoplankton groups in the region (diatoms and Phaeocystis antarcticd), macronutrients

(silicate, phosphate, and nitrate), the micronutrient iron (Fe), zooplankton, and detritus.

The Fe cycle used in this version of the model is simple and includes neither

photochemistry of Fe nor scavenging of Fe by particles, but still captures first order

dynamics of this essential micronutrient through remineralization, mixing, and advection

[Worthen andArrigo, 2003]. Remineralization of detritus is a source of dissolved Fe, and

there is weak restoring of Fe in the bottom layer of the model which simulates a

sedimentary flux of Fe. Phytoplankton growth rate is computed as a function of

temperature, light, and nutrient concentration.

Using climatological data sets as described by Arrigo et a/. [2003], the model was

spun up for three years and then run for one year to simulate the bloom over an average

annual cycle. The model is able to accurately simulate the temporal and spatial evolution

of two well understood phytoplankton blooms that occur annually in the Ross Sea

12

Page 26: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

polynya. The progression of the blooms and the magnitude of primary production are in

excellent agreement with both in situ observations and satellite ocean color data \Arrigo et

al, 2003; Worthen andArrigo, 2003].

2.3 Results

As described previously, the Chi distributions in the Ross Sea during the spring

phytoplankton bloom appear to be related to the bathymetry on the continental shelf, 300

- 600 m below the surface (Figure 2.2). The regions of low Chi (Figure 2.2e: lines 1

through 3) run parallel to the bathymetry contours of the troughs between the banks, but

are actually neither in the troughs nor on the shoals. Line 1 (Figure 2.1) runs down the

center of Mawson Bank and lines 2 and 3 border the largest shoal on the shelf, the

Pennell Bank.

Modeled surface velocities exhibit flow almost parallel to bathymetric contours,

both along the continental shelf break and on the shelf (Figure 2.3). Currents along the

170°E 180°

72°S

74°S

76°S

-1000 -800 -600 -400 -200 0

Figure 2.3 Instantaneous surface velocity streamlines overlain on bathymetry.

13

Page 27: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

shelf break move at approximately 0.1 m s"1, similar to flow estimates from previous

modeling studies \Assmann et al., 2003; Van Woert et al., 2003; Dinniman et al, 2003]. Our

results suggest that jets of oceanic water move onshore on both sides of the Pennell bank

and along the center of the Mawson Bank while offshore flow occurs along the canyon

between Mawson and Pennell banks. Although most previous modeling studies \Assmann

et al, 2003; Van Woert et al, 2003] were too coarse to resolve features in the circulation on

the shelf at a resolution suitable for the purpose of this investigation, our results agree

with circulation patterns derived from a higher resolution study \Dinniman et al., 2003]

showing onshore flow at 300 m depth flanking both sides of the Pennell Bank. Also

consistent with our findings is the observation that an influx of water from offshore may

be concentrated along the western flanks of the submarine banks on the shelf \Hofmann

andKJinck, 1998]. Current patterns predicted by CIAO are consistent with the few

measurements that are available, including long-term current meter observations of

northwest flow along the shelf break of 6.4 cm sA at 230m over a duration of 399 days

and one observation on the shelf at 230 m with velocities of <1 cm s"1 [Picco eta/., 1999 in

(a) 170°E 175°E 180°E (b) 170°E 175°E 180°E mw

Figure 2.4 Comparison of December chlorophyll distributions from (a) model results and (b) SeaWiFS imagery from 1998.

14

Page 28: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Dinniman etal., 2003]. Locamini [1994] used current meter measurements on the Ross Sea

continental shelf between 250 m and 450 m to identify a northward flow along the

western side of the shelf (near 170°E) as well as in the central shelf region (near 175°W),

in agreement with our model results. Unfortunately, their data are very sparse and

observations were not available near the shoals. A study of sea ice trajectories

(determined by tracking ice floes over time) show ice in the northwest area of the shelf

moving in a northward direction \Arrigo etal, 1998], consistent with our modeled

circulation patterns.

(a) (b) 170°E 175°E 180°E 170°E 175°E 180°E

33 33.2 33.4 33.6 33.8 34 -1 -0.5 0 0.5 1 1.5 2

Figure 2.5 Model results showing surface (a) salinity and (b) temperature and vertical profiles along a transect of 74.8°S of (c) salinity and (d) temperature

15

Page 29: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

The model predicts plumes of high and low-Chl waters (Figure 2.4a) that

correspond remarkably well to the high and low-Chl waters seen in satellite imagery

(Figure 2.4b). Although the horizontal extent and Chi concentrations of the modeled

bloom are greater than the bloom observed most years in the satellite imagery, the model

is able to capture the two areas of low Chi on the west side of the shelf (Figure 2.4a: lines

1 and 2). Surface salinity and temperature predicted by the model (Figure 2.5a and 2.b)

also exhibit alternating patterns of high and low values similar to that of Chi. Specifically,

areas of reduced salinity and temperature correspond to regions of low Chi. This pattern

was most pronounced in the salinity distributions (Figure 2.5a), which is not surprising

considering that density differences in polar waters are more strongly influenced by

salinity than by temperature. Vertical profiles of temperature and salinity (Figure 2.5c and

d) show that the horizontal patterns of these two properties in the surface plots are

restricted to water masses isolated above the pycnocline.

2.4 Patterns in Chi and Mechanisms for their Formation

Two mechanisms by which deep bathymetry can affect phytoplankton

distributions in surface waters, aside from facilitating the upwelling of nutrients, a process

not likely to be important in the high nutrient Ross Sea, include alteration of MLD and

modification of surface advection patterns. MLD can be altered when stratified waters

flow over irregular bottom topography. If the flow velocity over bottom topography is

subcritical (slower than the speed at which waves travel), the pycnocline will be displaced

upward and MLD will decrease. If the flow is supercritical (faster than the speed at

which waves travel), the pycnocline will be displaced downward and MLD will micken.

The location of the deepening or shoaling relative to the topography in each of these

16

Page 30: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

cases is impacted by local Coriolis strength. In areas where MLD has deepened,

phytoplankton are mixed out of the euphoric zone and bloom formation is inhibited. A

shallower MLD is often characterized by higher mean light levels, enhanced

phytoplankton growth, and higher Chi concentrations. Changes to MLD through this

mechanism can be substantial in a stratified oceanic environment where the interface

between the two internal layers of different densities can be depressed or raised to a

much greater extent than the sea surface [Stewart, 2004].

If this process were important in the Ross Sea, we would expect to see a shoaling

or deepening of the mixed layer associated with the stratified flow over the three large

banks on the continental shelf (Figure 2.1). Depth profiles of temperature and salinity

from CIAO (Figure 2.5c and 2.5d) reveal that the pycnocline remains relatively uniform

over the shoals, ranging from 10-30 m. Furthermore, waters on the Ross Sea continental

shelf are generally weakly stratified \Arrigo et a/., 1998; Arrigo et al, 2000], rendering it less

likely that stratified flow could alter the mixed layer enough to affect phytoplankton

growth. This indicates that stratified flow over the shoals is not markedly altering the

depth of the pycnocline to create the patterns in Chi, salinity, and temperature, either in

the model or in nature.

Model results suggest instead that the surface patterns in Chi are caused by

horizontal advection of different surface water masses, as revealed by vertical profiles of

temperature and salinity above the pycnocline (Figures 2.5c and 2.5d). Waters with low

Chi are associated with distinct water masses having low salinity and temperature (Figure

2.5) that advect in from northeast of the continental shelf. Modeled instantaneous

horizontal surface velocities (Figure 2.4a) give further indication that the patterns of high

17

Page 31: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

and low Chi are caused by horizontal advection. The surface velocities show that

onshore transport of oceanic waters are bathymetrically constrained to the troughs in

between the shoals (Figure 2.3). This pattern of modeled surface velocity persists

throughout December and January. Velocities along the shelf break are approximately 10

cm s"1, onshore velocities between the shoals are approximately 4 cm s"1, and offshore

velocities are approximately 2 cm s"1.

Chi concentrations in offshore waters that advect onto the shelf in association

with the jets are most likely low because these waters have not been ice-free long enough

for a bloom to form. As the temperature and salinity characteristics of waters within the

jets shows (relatively cold and fresh relative to surrounding waters, Figure 2.5), these jets

advected onto the shelf from regions that had recently been ice covered. Because the

lack of light in ice-covered waters reduces phytoplankton growth rates dramatically, these

newly ice-free source waters for the jets (such as along line 2 in Figure 2.4) have Chi

concentrations of only approximately 0.2 mg m"3. Given phytoplankton growth rates at

0°C of 0.2 d"1 in open waters of the Ross Sea, it would take approximately 30 days to

(a) 170°E 175°E 180°E (b) 170°E 175°E 180°E

Figure 2.6 Modeled results for surface waters in December of (a) chlorophyll and (b) iron.

18

Page 32: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

reach biomass levels characteristic of the surrounding high Chi waters within the bloom.

This is longer than the approximately 17 days that it takes for the jets to move from

offshore into the bloom. Thus, there is insufficient time for phytoplankton abundance

within the jet to reach concentrations characteristic of surrounding waters.

Low iron concentrations in the waters being entrained into the plume from

offshore may also play a role in maintaining the persistently low Chi concentrations

within the dominant jet (line 2). Results from CIAO show that iron concentrations in

December are low (below growth-limiting levels for phytoplankton) both within this jet

and in the offshore waters entrained into it (Figure 2.6: line 2). In situ iron

concentrations measured throughout the southwestern Ross Sea support the idea of iron

limitation within offshore source waters for the jets. Although iron abundance in surface

waters on the Ross Sea continental shelf before the spring bloom is relatively high (0.1 -

0.4 nM), concentrations decrease markedly northward, falling to 0.025 nM (in mid­

summer) at the edge of the continental shelf and to <0.05 nM (year-round) in the

Antarctic Circumpolar Current [Coak et al, 2005]. However, it should be noted that iron

concentrations in seawater and particularly in ice and snow are still not well known and

have been parameterized in CIAO based on only a limited number of samples from the

Ross Sea [Coak et al, 2005; Edwards and Sedmck, 2001]. More in situ data are needed to

firmly establish surface Fe concentrations in the Ross Sea and to ensure that iron

concentrations are being properly represented in non-shelf waters within CIAO.

If the areas of reduced Chi concentration within the phytoplankton bloom result

from the advection of offshore waters low in Chi into the bloom, this may explain why

the easternmost jet of low-Chl water (Figure 2.2: line 3) is less pronounced during years

19

Page 33: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

when the phytoplankton bloom is more extensive (Figures 2.2a-c) and why it is not well

represented by CIAO (Figure 2.4a: line 3). Unlike the dominant low-Chl jet (line 2.2),

which is formed from the southwestward flow of mostly non-shelf waters onto the

continental shelf (Figure 2.3), much of water forming the easternmost jet of low Chi

water originates on the shelf, flows northward toward the shelf break, then turns back

toward the southwest before forming the jet and into the bloom (Figure 2.3: line 3).

During years when the bloom is more extensive (e.g. 1978 and to a lesser extent, 1998,

Figures 2.2a and 2.2b), waters on the shelf where the jet originates will have a high Chi

content. Consequently, although this northward flow entrains some low-Chl offshore

water before moving back onto the shelf, the jet is much less pronounced than in years

when the bloom on the shelf is less extensive (Figure 2.2c).

2.5 Advection via Ekman Layer, Tidal, and Geostrophic Flow

Having determined that horizontal advection controls phytoplankton

distributions in the Ross Sea, we next consider which of a number of different advective

processes (surface Ekman dynamics, tides, and geostrophy) could respond to bathymetric

features hundreds of meters below and may be responsible for the patterns of advection

seen in surface waters. On low-latitude continental shelves, phytoplankton blooms are

often affected by the relatively shallow bathymetry through turbulent mixing to the sea

floor [Lucas and Cloern, 2002] or topographically induced eddies bringing nutrients into the

euphotic zone [Roman eta/., 2005]. However, Antarctic continental shelves are deep

compared to low latitude shelves (typically hundred of meters versus tens of meters

respectively), and these processes are likely to be of little importance.

20

Page 34: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Surface circulation is most directly affected by wind stress creating flow in die

surface Ekman layer, but this type of flow is unlikely to form consistent patterns in

surface circulation related to bathymetry on the Ross Sea continental shelf. Winds

blowing on the sea surface produce a thin boundary layer which is at most a few-hundred

m thick, but usually only 5-30 m thick [Stewart, 2004]. Surface Ekman layer flow does not

extend to the ocean floor in the Ross Sea, and cannot connect surface circulation to deep

bathymetric features.

Tidal models \Padman et al, 2003] predict that tidal currents are generally strong in

the southwestern Ross Sea, particularly over the shoals on the continental shelf.

Furthermore, tidally driven flow is spatially consistent over these features from year to

year. Nevertheless, model results rule out tides as a first-order factor in the formation of

these jets and plumes, as these features appear in the model results when tidal forcing is

excluded. Additionally, tidal current ellipses [MacAyeal, 1984] are on the order of 10 to

100 times smaller than the mesoscale patterns observed in the phytoplankton bloom.

There is no obvious mechanism by which tidal flow, circling around in these small-scale

tidal ellipses, could create the observed patterns in the bloom through advection alone.

We conclude then that the plumes of water observed in the model output and in

the satellite imagery are caused by geostrophic flow, whereby fluid motions are

maintained as a result of near balance between the Coriolis effect and gravity-induced

horizontal pressure gradient forces. This conclusion is based on the previously described

relationship between surface advection and the shoals on the continental shelf, as well as

velocity streamlines that are nearly parallel to isopycnal contours (Figure 2.7), a

characteristic of flow in geostrophic balance. Furthermore, polar regions experience

21

Page 35: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

170°E 175°E 180°E

Figure 2.7 Instantaneous surface velocity streamlines overlain on modeled surface density anomaly.

strong rotation due to their high latitudes and are strongly influenced by geostrophic

flow. Consistent with this, currents in the Ross Sea study region are dominated by

geostrophic flow, with well documented evidence of this flow apparent in geologic

features from the Ross Sea upper slope [Anderson eta/., 1984]. This type of flow extends

deep into the water column and tends to move around obstacles such as shoals, instead

of over them, connecting deep bathymetric features to surface waters above.

2.6 Conclusions

Previous studies provide evidence that annual production in the southwestern

Ross Sea is limited by iron availability [Cochlan et a/., 2002; Olson et a/., 2000; Arrigo et a/.,

2003]. In the present investigation we extend these studies to show that the horizontal

distribution of high productivity water is most likely constrained by geostrophic surface

currents that continually move waters low in Chi, and most likely, iron, both along the

shelf break and onto the shelf. Interannual variation in the size of the maximum

22

Page 36: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

phytoplankton bloom in the Ross Sea during years of normal sea ice cover may be the

result of annual differences in the balance between the strength of geostrophic circulation

along the shelf break, and upward mixing of iron from deeper waters on the shelf.

Reduced flow within die southwestern leg of the Ross Sea gyre or more intense vertical

mixing in winter in some years may allow the phytoplankton bloom to extend further

toward the continental shelf break.

This study also gives some insight into questions regarding the importance of

Circumpolar Deep Water (CDW) moving onto the Ross Sea continental shelf as a source

of nutrients for the spring phytoplankton bloom. Studies on the west Antarctic Peninsula

continental shelf have shown that topographically induced upwelling at shallow locations

near the shelf break brings nutrient-rich CDW into the euphoric zone and increases

biological activity [Pre^e/in eta/., 2000]. Dinniman etal. [2003] have suggested that similar

processes may be important in the Ross Sea. Warm CDW upwelling onto the Ross Sea

continental shelf has been proposed to contribute to the formation of the seasonal

polynya [Jacobs and Comiso, 1989], but there remains litde evidence that this water mass is

entrained into surface waters during the summer or is biologically significant. In fact,

CIAO results presented here suggest that on die shelf and across the shelf break, there is

a reorganization of surface flow as opposed to the entrainment of intermediate waters

into the surface. While prior studies suggested that deep offshore water enhances

biological activity in the Ross Sea polynya, we suggest alternately that in some regions, the

offshore surface waters moving onto the shelf are inhibiting phytoplankton growth.

23

Page 37: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 3 The Role of Thermal and Mechanical Processes in the Formation of the Ross Sea Summer Polynya

Abstract

Three decades of satellite observations collected during spring and early summer

have shown a recurring region of ice-free water forming in the sea-ice cover of the Ross Sea,

Antarctica. This Ross Sea summer polynya plays an important role in heat exchange between

the ocean and atmosphere, ventilation of deep water, and is characterized by high biological

productivity. Despite its appearance each year, the relative importance of different physical

processes to its formation and maintenance are not widely agreed upon. Here we use a three-

dimensional coupled ice/ocean model to better understand processes controlling the

dynamics of the Ross Sea polynya. Results from the model control run agree favorably with

satellite microwave imagery of sea ice. Model sensitivity studies suggest that polynya

dynamics are insensitive to the amount of snowfall, the presence of the Ross Ice Shelf cavity,

tides, and solar radiation penetrating the ice. The model results also corroborate earlier

findings that both the advection of sea ice and heat entrainment from warm Modified

Circumpolar Deep Water play a role in Ross Sea polynya development. More importantly,

the model further demonstrates that advection of sea ice due to wind stress plays the primary

role in summer polynya formation. Additionally, we suggest that 1) heat entrainment reduces

the rate of sea ice formation rather than melts existing sea ice, and 2) advection of sea ice due

to synoptic wind events associated with variations in atmospheric pressure are the processes

primarily responsible for the formation and expansion of the Ross Sea summer polynya.

24

Page 38: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

3.1 Introduction

Because polynyas are areas of open water in regions otherwise covered by sea ice,

they have profound influences on the physics, chemistry, and biology of the waters in

which they are formed. They are sites of enhanced exchange of both heat [Dare and

Atkinson, 2000] and gases [Sweeney et al, 2000] between the ocean and the atmosphere.

While ocean-atmosphere heat transfer is especially high in winter due to the large air-sea

temperature gradient and higher wind speeds [Roberts et al, 2001], exchange of gases such

as carbon dioxide, oxygen, and dimethylsulfide are greatest in summer and autumn due to

enhanced biological activity [Arrigo et al, 1999; 'bates et al, 1998]. Due to the increased

availability of light in nutrient-rich waters not covered by ice, polynyas are often sites of

high primary and secondary production [Arrigo and Van Dijken, 2003; Reddy and Arrigo,

2006] which supports large populations of fish, benthic communities, mammals, and

birds [Ackley et al, 2003; Ainky et al, 1978; Ainley et al, 2005; Aoyanagi and Ueda, 1989;

Clarke and Johnston, 2003; Court et al, 1997; Dewitt, 1970; Erickson et al, 1971; Polito et al,

2002]. Additionally, the brine rejected as sea ice freezes in many coastal polynyas affects

the density of regional water masses [Bindoffet al, 2001; Buffoni et al, 2002; Zwally et al,

1985], vertical mixing, and modifies circulation on a global scale [Gordon and Comiso, 1988;

Grigg andHolbrook, 2001]. Although it varies interannually in extent and shape, the Ross

Sea polynya is the largest (~400,000 km2 in the summer) and most biologically productive

(100-200 g C m 2 yr4) polynya in the world [Arrigo and Van Dijken, 2003].

Past field and modeling studies of polynya dynamics \Bromwich and Kurt^ 1984;

Bromwich et al, 1993; Darby et al, 1995; Gallee, 1997; Gordon et al, 2000; Pease, 1987; Van

Woert, 1999; Wu et al, 2003; Bromwich et al, 1998; 1997] have focused primarily on winter

25

Page 39: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

polynyas because their impact on ocean circulation and ocean-atmosphere heat exchange

is greatest at this time of year and because environmental conditions in the winter are

simpler to model (shortwave radiation does not need to be accounted for in winter

months) \Bromwich eta/., 2001]. However, the importance of the springtime expansion of

these polynyas, driven by winds and open water-albedo feedback during periods of

increased solar radiation \Hunke andAckley, 2001; Ohshima andNihashi, 2005], are recently

receiving more attention \Brommch eta/., 2001] due to their large surface area for gas and

heat exchange and greater light availability for phytoplankton production. In this study,

we will refer to the larger area of open water in the spring and summer as a summer

polynya \Arrigo and Van Dijken, 2003; Maqueda eta/., 2004; Smith and Gordon, 1997;

Bromwich eta/., 2001].

Polynyas are formed by thermal and mechanical processes and have been

categorized as either sensible heat or latent heat polynyas, respectively. Thermally-driven

polynyas form when oceanic heat flux is sufficient to melt existing sea ice, prevent

formation of new ice, or slow the rate of ice accretion so that wind-driven advection or

increased solar radiation can more easily reduce local sea ice concentrations. Latent heat

polynyas generally form in areas where winds blow sea ice offshore to produce a region

of open water near the coast, with the maximum polynya size being a function of air

temperature [Pease, 1987] and wind speed. New sea ice continually forms in this open

water, releasing latent heat as it freezes, and is subsequently blown offshore. Thus, latent

heat polynyas are not actually formed by the release of latent heat, but rather, are formed

due to mechanical processes such as katabatic or synoptic winds moving sea ice out of a

region [Maqueda eta/., 2004; Bromwich et a/., 1998]. Due to the confusing nomenclature, we

26

Page 40: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

will refer to polynyas here as being either thermally- or mechanically-driven. It should be

noted, however, that most polynyas are thought to be formed by a combination of both

thermal and mechanical processes [Smith etal., 1990].

There has been debate in the literature about the relative role that thermal and

mechanical processes play in the rapid springtime expansion of the Ross Sea polynya.

Although it has been considered primarily to be a wind-driven polynya \Zwally et al,

1985], the Ross Sea polynya may be, at least in part, thermally driven {Jacobs and Comiso,

1989]. The likely oceanic heat source is the relatively warm Modified Circumpolar Deep

Water (MCDW) that is transported southward onto the continental shelf and is entrained

into the surface layer during winter mixing \Dinniman et al, 2003; Gordon et al, 2000; Arrigo

and Van Dijken, 2004; Jacobs and Comiso, 1989]. This additional heat reduces the rate of sea

ice growth, resulting in an ice pack with reduced sea ice concentrations and thickness,

thereby facilitating polynya formation. However, because most previous studies have

focused on the Ross Sea polynya during winter months, less is known about the factors

that drive the formation of the summer polynya. To assess the relative roles of thermal

and mechanical processes in the formation and maintenance of the Ross Sea summer

polynya, we present results from a three-dimensional regional modeling study focusing on

the expansion of the polynya during the spring and summer months. We used a coupled

sea-ice/ocean model to investigate the relative role of winds, sensible heat flux, ocean

currents, the Ross Ice Shelf (RIS) cavity, tides, snowfall, and solar radiation in the

formation and maintenance of the polynya. By comparing results from the numerical

model to satellite-derived sea-ice concentrations and field measurements of ice and snow

thickness, we were able to assess the factors that drive the regular formation of the Ross

27

Page 41: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Sea polynya and determine the relative contributions by mechanical and thermal

processes.

3.2 Methods

3.2.1 Model Description

The Ross Sea polynya was modeled using the Polar Ocean Land Atmosphere and

Ice Regional (POLAIR) modeling system {Holland and Jenkins, 2001; Holland et al., 2003].

The model contains components representing ocean circulation, land margins (including

coastlines and rivers, if applicable), a simple atmosphere (which calculates radiation fluxes

and specifies temperature and pressure from observational data), and sea-ice

thermodynamics-dynamics. The ocean circulation component of the model is an

isopycnic, primitive equation code modified from the Miami Isopycnic Coordinate Ocean

Model (MICOM) [Bleck and Smith, 1990]. The ocean model includes parameterizations

that allow for strong lateral mixing along isopycnic layers as well as weak diapycnic

mixing between vertically adjacent isopycnic layers. These parameterizations mimic, to a

large extent, the observed mixing behavior in the real ocean. The sea-ice component uses

a two-thickness category approach, one category consisting of sea ice of a mean thickness

and the other representing leads between ice floes. Treatment of ice-ice interactions is via

a cavitating-fluid rheology \Flato and Hibler, 1992] in which the ice pack is resistant to

convergence but not divergence. Sea-ice concentration and thickness are computed by

accounting for the divergence of heat and freshwater fluxes at the interfaces between the

sea-ice and other model components (ocean and atmosphere). In many ways, the sea ice

is modeled as an "isopycnic" fluid layer, just as the ocean is, but with different stress

properties [Holland and Jenkins, 2001]. The equations governing sea ice are as follows:

28

Page 42: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

^ + V ^ + ( / + 6)*xv /+VM, = {Tia r*>+lv<<frVv,) (1)

dt ^ + V.(vA) = V . « V ^ ) + 3/" (2)

V.(v/C,) = V.(* /cVC /) + 3f (3)

a/ where v; is the horizontal velocity vector, / is the Coriolis parameter, 4 is the relative

vorticity, "• is the vertical unit vector, Mt is the Montgomery potential, xia is the wind

induced sea ice surface shear stress vector, xoi is the analogous sea ice bottom-drag stress

vector, PJ is the density of the ice, ht is ice thickness, ' is the eddy viscosity, ' and '

are the eddy diffusivities, Cf is the concentration and ' and ' are the sum of all

thermodynamic forcing on the thickness and concentrations fields, respectively. See

Holland et al. [1993] and Holland [1998] for a complete discussion of ice thermodynamics.

Each grid cell has a computed ice concentration which is defined as the fraction (0 to 1)

of the grid cell which covered by sea ice. The remaining fraction of the grid cell is

defined to be the areal fraction of leads (open water). A viscous-sublayer

parameterization [Holland and Jenkins, 1999] describes thermodynamic interactions at the

interface between the ice shelf base and ocean surface as well at the sea-ice - ocean

interface. The ocean model has been modified to incorporate the arbitrary surface

topography of an ice shelf base [Holland and Jenkins, 2001]. The ocean cavity below the

RIS is modeled as described by Holland et al. [2003]. Tides are simulated using local non-

equilibrium tidal forcing generated by the Circum-Antarctic Tidal Simulation (CATS)

model {Robertson et al, 1998]. Snow accumulates on the sea ice and visible solar radiation

penetrates through snow and ice according to Beer's law such that:

29

Page 43: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

HT ^ ^-d_icezice 0 ^-d_snowzsnow /r\

where T is the fraction of incident light that is transmitted through the snow and ice, KdJa

and jK^^are the attenuation coefficients for ice (1.4 m4) and for snow (12 m"1),

respectively, and %f«and Zm» are the thicknesses of the ice and snow, respectively, in each

grid cell. The model grid is based on a Mercator projection and all grid cells are rendered

to be locally square. Moving southward in the model domain, grid cells become smaller

in proportion to the cosine of the latitude. For the control model run and most

experimental runs, the average horizontal grid box size in the middle of the Ross Sea

continental shelf is approximately 70 km. The model domain is comprised of 40 grid

cells in the longitudinal direction from 140°E to 120°W and 33 grid cells in the latitudinal

direction from 60°S to 85°S (Figure 3.1). This domain includes the Ross Sea continental

shelf, the deep ocean, as well as the cavity beneath the RIS (Figure 3.2).

Figure 3.1 Map of the Antarctic coastline and the 1000 m bathymetric contour. The Ross Sea model domain perimeter is shown (heavy outline) as well as the perimeter for the model domain which excludes the Ross Ice Shelf (dashed line).

30

Page 44: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

60°S

66°S

<u •u 3 72°S

78°S

84°S

140°E 160°E 180" 160°W 140°W 120°W

-4000 -3000 -2000 -1000 0

Figure 3.2 Bathymetry (m) of the Ross Sea model domain which includes the Ross Ice Shelf cavity, the continental shelf and the abyssal ocean. Also shown are the three regions for which statistics were extracted.

Datasets defining the surface of the sub-ice shelf cavity, sea floor, and initial

density structure of the model ocean are described by Holland et al. [2003]. NCEP

reanalysis products [Kahay et al, 1996] for air temperature and sea-level pressure are used

as atmospheric forcing. The model computes geostrophic winds by computing the

horizontal gradient of daily sea-level pressure. In the absence of a robust dataset for

humidity, a value of 0.90 was used by the model to capture relative humidity over the

ocean. In lieu of a dataset quantifying cloud cover over sea ice, a constant cloud cover of

0.85 was used for the region, based on average cloud cover from climatology. Snow

thickness is derived from a global precipitation climatology [Legates and Willmott, 1990]

where precipitation over sea ice increases snow thickness based on the density of snow

(300 kg m3). Snow thickness can decrease through melting, sublimation, or when snow is

transformed into sea ice. The latter occurs when there is sufficient snow cover to

submerge the sea-ice freeboard.

31

Page 45: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

A sponge zone of three grid cells is used for the ocean model to restore

temperature and salinity to a prescribed pattern based on climatology along the northern,

eastern, and western sidewalls of the open ocean part of the domain. The exterior-most

sponge cells are restored on a 10-day timescale and the interior-most on a 30-day

timescale for baroclinic fields, and on 1-day and 5-day timescales, respectively, for

barotropic fields. This approach provides reasonable lateral boundary conditions, helps

prevent artificial drift due to the model's limited domain, and ensures that properties at

the boundaries are kept close to observed values. The model was spun-up by cycling the

same year (1 January through 31 December 2000) of forcing data for 10.5 years.

3.2.2 Control Run and Parameter Sensitivity

For the control run, we included all processes believed to have the potential to

influence polynya size and shape, including sea-ice dynamics and thermodynamics, the

cavity under the RIS, solar radiation penetrating through sea ice, tides, precipitation, and

daily wind forcing. We also ran the model without restoring sea-surface salinity to

observed data allowing the surface salinity to evolve freely. The model was run at 2.5°

horizontal resolution for 10.5 years, with 11 vertical layers in the ocean component. A

number of parameters for which little is known for the Ross Sea were tuned based on

ranges of values obtained from the literature to give the most accurate sea-ice simulation

compared to satellite data from SSM/I. The tuned parameters (summarized in Table 3.1)

include the coefficient of drag between the ocean and ice [McPhee, 1983; Overland and

Davidson, 1992; Overland et al, 1984], the planetary boundary layer (PBL) turning angle (the

offset in the angle of the wind at the lower boundary of the atmosphere due to Ekman

layer flow with an uncertainty of approximately 10°) [McPhee, 1980], the albedo of the sea

32

Page 46: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Table 3.1 Uncertain model parameters and coefficients used in sensitivity analyses

PBL turning angle [degrees] ocean-ice drag coefficient [dimensionless] ice albedo [dimensionless] ocean albedo [dimensionless] ice freezing coefficient [dimensionless] ice melting coefficient [dimensionless]

Control Run 25

6.0E-3 0.60 0.06 4.0 1.5

Sensitivity Run Low Value

15 3.0E-3

0.50 0.04 2.0 0.75

Sensitivity Run High Value

35 9.0E-3

0.70 0.10 6.0 2.0

ice and the ocean, and the melting and freezing coefficients for sea ice (see below). The

values in the literature for parameters such as the melting and freezing coefficients for sea

ice were experimentally derived. Within this range, the values for each parameter used in

the control run were narrowed to a single number by tuning. Thus, they are validated

model assumptions, the range for which were experimentally derived. Separate runs were

performed to assess the sensitivity of model results to the choice of coefficients for each

of these parameters.

The melting and freezing coefficients for sea ice determine the rates at which ice

melts and freezes in horizontal versus vertical directions [Holland, 1998]. The

concentration of ice is determined by the coefficient of freezing such that

c,+1 = c, + o-c,) coeft . S.7.. ^ V freezing^^ ice (5)

''lead

where C is the concentration of ice, the subscript t indicates the time index for

discretization, coeffree2ing is the coefficient for freezing, Azice is the change in ice thickness

and zlead is the thickness of ice in leads through which ice grows and equals 0.5 as given by

Hibler [1979]. Similarly, the melting of ice is determined as

ci+1=ct+(o.5Cty coef u- Az.

^ / melting ^^ ice maX(Zice'Zice__mm)

where coefmeWng is the coefficient for melting, zice is the thickness of the sea ice and zic

is the minimum allowable thickness of the modeled sea ice.

(6)

33

Page 47: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Table 3.2 Model configurations for sensitivity runs

Model Configuration Description (a) Control run Heat is not entrained into the mixed layer from deeper layers from (b) June — March (c) December -March (d) the entire 10 year spin-up period (e) Sea ice does not move from one place to another but still undergoes thermodynamical processes (June - March) (f) Heat is not entrained into the mixed layer from deeper layers and sea ice does not move from one place to another but still undergoes thermodynamical processes (June — March) (g) The open water albedo feedback effect is examined by setting the albedo of the ocean to 1 (June - March) Wind forcing is not included in the model (h) June - March (i) December -March Snow is removed and precipitation is turned off resulting in no associated freshwater inputs to the ocean from snowfall (i.e. there are still freshwater input from ice melt) and no accumulation of snow on die sea ice (j) June -March (k) December - March (1) The cavity under the RIS is not included in the model (m) The model is run without tidal forcing (n) Solar radiation is not transmitted through sea ice (o) Daily solar radiation is calculated at hourly time steps instead of using a daily average (p) Sea-surface salinity is restored to climatological values (q) The ocean component was run with 25 instead of 11 vertical layers

3.2.3 Model Configuration Sensitivity Runs

To better understand the factors that affect the formation of the summer polynya

in the Ross Sea, we also performed a number of model configuration sensitivity

experiments. Because the control run was the most realistic model setup, sensitivity runs

were executed by excluding or modifying one model component at a time (summarized in

Table 3.2) to determine its effect on modeled polynya formation. The effect of excluding

the cavity below the RIS (run 1, Table 3.2) was examined by moving the southern model

boundary from 85°S to 80°S and making it a closed wall. It was determined that

excluding the cavity had virtually no impact on the modeled polynya formation. Thus, to

save computational expense, all subsequent model configuration sensitivity runs were

conducted with a domain that excluded the RIS cavity (and run k became the control run

34

Page 48: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

for the model configuration sensitivity study). The sensitivity runs were executed for one

year with the change imposed on 1 June (unless otherwise specified) after model spin up.

Runs that excluded the entrainment of heat (runs b-d, Table 3.2) allowed water volume

but not heat to be entrained from deeper layers of the ocean into the mixed layer. To

exclude ice dynamics (run e, Table 3.2), sea ice was not allowed to move from one grid

cell to the next, but could still undergo thermodynamical processes. Another run

excluded both the entrainment of heat as well as ice dynamics (run f, Table 3.2). The

effect of open water albedo feedback was examined by setting the albedo of the ocean to

1 (run g, Table 3.2). Sensitivity to winds was evaluated by setting the wind speed to zero

over different time periods (runs h and i, Table 3.2). Similarly, snowfall was eliminated by

setting precipitation to zero over the same two periods, which eliminated both freshwater

input to the ocean surface from snowfall as well as snow accumulation on the sea ice

(runs j and k, Table 3.2). Tides were excluded from the model by eliminating tidal forcing

from the CATS model (run m, Table 3.2). The run without solar radiation penetrating

through sea ice (run n, Table 3.2) warmed the ice but did not warm ocean waters beneath

the ice.

Another set of sensitivity runs incorporated model setups that were more realistic

than the control run, but were either more computationally expensive or overly

prescribed. For example, in one run we calculated solar radiation at hourly time steps

instead of using a daily mean (run o, Table 3.2). An additional run was performed (run p,

Table 3.2) in which ocean surface salinity was restored to climatological values (with a

restoring timescale of 30 days) using a pattern of salinity based on summer observations

(as described by Holland et al. [2003]). To assess the effect of resolution on model

35

Page 49: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

predictions, we increased the number of vertical layers in the ocean from 11 to 25 (run q,

Table 3.2). Because of computational expense, lower vertical resolution runs were used

for the control run and to perform all other sensitivity experiments.

Results from sensitivity experiments were compared to those of the control run

results for the months of January, February, and March. To quantitatively compare

model results from the control run and sensitivity runs, statistics were extracted from

three geographical regions (Figure 3.2) having distinct ice dynamics, including the western

Ross Sea (Western, 64°S-68°S and 145°E-160°W), the Ross Sea polynya (Polynya, 70°S-

78°S and 160°E-170°W), and the eastern Ross Sea (Eastern, 70°S to 74°S and 140°W to

125°W). These three regions were chosen for statistical analysis because ice often flows

from the eastern Ross Sea to the polynya region and then northwest to the western Ross

Sea due to the prevailing winds and ocean currents. Analyzing the three independently

Table 3.3. Model sensitivity analyses Run

a b

c

d

e

f

g

h

i

i k

Model Configurations

Control run No heat entrainment from June - March

No heat entrainment from December-March No heat entrainment for the entire 10 year spin-up period Ice not moving from June - March

No heat entrainment and ice not moving from June - March Open water albedo feedback effect from June - March No wind from June - March

No wind from December - March

No snow from June - March

No snow from December - March

Fractional Ice Cover 1 January - 1 March and percentage differences compared to control run

Western 0.725 0.954 32% 0.633 -13% 0.986 36% 0.00

-100% 0.515 -29% 0.684 -6%

0.920 27% 0.823 14% 0.391 -46% 0.445 -39%

Polynya 0.494 0.737 49% 0.468 - 5 %

0.953 93% 0.760 54% 0.993 100% 0.528 7%

0.960 94% 0.793 61% 0.495 0%

0.471 -5%

Eastern 0.919 0.703 -24% 0.924 1%

0.935 2%

0.937 2%

0.718 -22% 0.872 -5%

0.807 -12% 0.822 -11% 0.724 -21% 0.764 -17%

36

Page 50: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

provided a better overall view of sea-ice dynamics over the entire domain. Ice

concentration in each region was averaged from January through March (Table 3.3). In

addition to computing average sea-ice concentration in each region from January through

March, the change in ice concentration relative to the control run was also calculated

for each. This was computed as the difference in ice concentration between the

experimental run and the control run, normalized by the ice concentration of the control

run. Thus,

C -C n ice_sensitivity ice_control . /~,^ /r-f\

~ c ice _control

where R the percent change in ice concentration relative to the control run, Cia smjjtjvjty is

the mean ice concentration for a given region in the sensitivity run, and C^^^, is the

mean ice concentration for a given region in the control run.

3.3 Results

3.3.1 Control Run

Our model control run simulated the summer formation of the Ross Sea polynya well,

not agreeing perfectly, but capturing the basic physics of the system (Figure 3.3). In

December, the maximum extent of the modeled sea ice is virtually identical to satellite-

based measurements, with ice extending northward to between 62°S and 66°S. The

model also accurately captures the polynya opening north of the RIS on the western side

of the continental shelf in December and expanding to the full width of the ice shelf

front in January and February. Satellite data show that by January, the Ross Sea summer

polynya has broken through the remaining strip of sea ice separating the polynya from the

open ocean to the north, forming a single contiguous body of ice-free water (Figure 3.3).

37

Page 51: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

jQ

E

o

-Q

E <D O CD Q

!

2 * 3 .Q

2 CO

2

CL <

140°E 160°E 180° 160°W 140°W 120°W 140°E 160°E 180° 160°W 140"W 120°W

Longitude Longitude

20 40 60 80 100

Figure 3.3 Sea ice concentrations from (a) satellite climatology (1987-1993) and (b) year 10 of the model control run

38

Page 52: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

The model captures this pattern, with a thin band of open water and greatly reduced sea-

ice concentrations in the breakthrough region, although this breakthrough occurs

approximately a month late in the model. Furthermore, the simulated Ross Sea summer

polynya reaches its maximum size in January/February (Figure 3.3), similar to

observations. The refreezing of the summer polynya is captured particularly well by the

model, with the summer polynya beginning to close in March in both the satellite

observations and the model due to rapid sea-ice production just north of the RIS (Figure

3.3). By April, both the modeled and observed polynyas are almost fully closed, and the

northern extent of the ice over most of the domain reaches 68-70°S, although the extent

in the model simulation is not as far north as in observations (Figure 3.3). The largest

difference between observed sea ice dynamics and the modeled Ross Sea polynya is the

northern extent of the ice during January, February, and March in the western sides of the

60°sJ

c 3

- J

60°S

70°S

a

;«?* 3 § & ^ l

- j . . - i _ _ i.. i. ,"1-

1 | b - !

i

^"^T^^^TI^H ''.JipR

, . . J .J ..i

, a g ^

L^stggifF' -

W^^B**'̂ —

(0 3 u

.a u. to 3

60°S

70°S

\ d

\

A" •' ..-"'* ' ' • v .

„ • , ; >

*»4

\J •Jl

.,, }

^3 ' **fi

*y-A

/

, *•

w W :' r*%,

140°E 180° 140°W Longitude

140°E 180° 140°W Longitude

-mm [m]

140°E 180° Longitude

140°W

0 0.1 0.3 0.7 1.2 2.0 0 0.5 1.0 1.5 2.0 2.5 3 [m]

Figure 3.4 Sea-ice thicknesses from (a) June 1995 data (b) June 1998 data (c) June 1st model control run (d) January 1999 data (e) January 2000 data (f) Feb 1st model control run (sea-ice thickness data compiled by DeUberty and Geiger [2005])

39

Page 53: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

domain (Figure 3.3). The model shows relatively dense multi-year ice in this region,

whereas observations show the ice concentrations are somewhat lower.

Modeled ice thicknesses are consistent with two seasons of sea-ice thickness data

from satellite-derived National Ice Center charts for June and January and compiled by

DeUberty and Geiger [2005] (Figure 3.4). In June, both the data (Figures 3.4a and b) and

the model results (Figure 3.4c) for the western side of the domain (near 160°E) show sea

ice decreasing in thickness from 2 m near shore to approximately 0.1 m offshore, although

on average, ice in the model was somewhat thicker than observations. In the Ross Sea

polynya region in June, both observed and modeled sea-ice thickness exhibit a distinct

east-west gradient, with markedly thicker ice being concentrated towards the east. The

primary difference is that the model produces maximum ice thicknesses of 3 m whereas

thicknesses from the ice charts are generally 2 m or less. In June, on the far eastern side

of the domain, ice thickness in both the model and the observations reaches its maximum

within a band located about 100-200 km offshore, decreasing in thickness rapidly to the

north. Again, the maximum ice thicknesses produced by the model are somewhat thicker

(2-3 m) than the observations.

In January, observed sea-ice thickness within the region of persistent ice in the

western portion of the model domain ranges from 0.7 m to 2 m (Figures 4d and e),

similar to model results of 1-2 m (Figure 3.4f). The polynya region is ice-free in January

in both the observations and model results, with the shape of the Ross Sea polynya in the

model corresponding quite well with observations from 1999 (Figure 3.4d). Finally, the

area of thick ice on the eastern side of the continental shelf corresponds well with

40

Page 54: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

observations, particularly in 1999, although the model produces ice of greater thickness

than observations.

We also compared modeled ice and snow thickness to field data collected along

two north-south transects in the Ross Sea polynya region, one from austral autumn

[Jeffries andAdolphs, 1997] and the other from austral spring \Arrigo et a!., 2003] (Figure

3.5). In May-June (Figures 3.5a and b), the observed mean ice thickness was 0.21 m near

the continent, increased to 0.65 m at a latitude of 76°S, and then decreased to 0.35 m

along the northern ice edge near 66°S. Modeled sea ice thickness exhibited a similar

latitudinal pattern, although sea ice was generally thicker and the region of maximum ice

thickness was located farther to the north than the observations. Observed snow

w w <D c o 1c I -

1.6-1.2-

0.8-0.4-00-

a) Autumn Field Data

^F*=3^-Y- • ,

Lb-1.2-

0.8-0.4-00-

b) Autumn Model Results

IK-ctLl-L_

78°S 76°S 74°S 72°S 70°S 68°S 66°S

c) Spring Field Data

78°S 76°S 74°S 72°S 70°S 68°S 66°S

d) Spring Model Results

78°S 76°S 74°S 72°S 70°S 68°S 66°S Latitude

78°S 76°S 74°S 72°S 70°S 68°S 66°S Latitude

Mmtmm,a Snow thickness i ' Sea-ice thickness

Figure 3.5 Comparison of ice and snow thicknesses (a) from May/June 1995 field observations along transects between 160°W and 180° {Jeffries andAdolphs, 1997] and (b) from June of the 10th year of the model control run along a 180° transect (c) November 1998 field observations along a transect of approximately 175°E [Arrigo etal., 2003] and (d) from November of the 10th year of the model control run along a 175°E transect

41

Page 55: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

thicknesses in May-June varied from 0.03 m to 0.15 m, averaging 0.10 m. Modeled snow

thicknesses were slightly greater than observed (0.14 m to 0.33 m) but exhibited a

comparable spatial pattern, with lowest accumulations in the far northern and southern

regions of the model domain. It should be noted, however, that observations of snow

and ice thickness are often of low spatial resolution and therefore are not necessarily

representative of true distributions due to high spatial variability in situ.

In November, observed ice thicknesses were around 0.1 m close to the continent,

grew to approximately 0.6 m at 72°S, remained approximately the same thickness

between 72°S and 66°S, and decreased to 0.5 m at the northern ice edge \Arrigo et a/.,

2003]. Modeled ice in spring was thicker than observed, but exhibited a similar spatial

pattern, with the thinnest ice being close to the continent and the thickest ice located in a

broad band centered at approximately 70°S. Observed snow thicknesses were very thin

in the southern reaches of the Ross Sea polynya close to the continent (<0.05 m) and

reached a maximum thickness of 0.2 m at 68°S. Modeled snow thicknesses in November

were also thicker than observed, ranging between 0.10 m (at 77°S) and 0.45 m (at 67°S)

and averaging approximately 0.25 m.

Although the model has a tendency to produce sea ice that is somewhat thicker

than has been observed in the Ross Sea, it reproduces the temporal ice kinematics with

remarkable fidelity, as well as the timing of polynya expansion and the spatial patterns of

sea-ice extent. The concordance with observations suggests that the first-order processes

responsible for the formation, expansion, and maintenance of the Ross Sea summer

polynya are well represented in the model realization used to produce the control run.

Thus, we were able to conduct sequential sensitivity studies to determine what aspects of

42

Page 56: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

the coupled ice-ocean-atmosphere polynya system are primarily responsible for the

observed kinematics of the Ross Sea polynya.

3.3.2 Sensitivity Studies

3.3.2.1 Sensitivity to Parameters Values

An initial set of experiments was performed to test the sensitivity of the model to

those parameters about which little is known for the Ross Sea, including the PBL turning

angle, the drag coefficient between the ice and ocean, the albedo of the sea ice and the

ocean, and the melting and freezing coefficients for sea ice. These sensitivity studies

showed that varying any of these parameters within the range of values from the literature

(as described in Table 3.1) altered sea-ice concentrations by only 1-4% relative to the

control run, and had litde impact on the size of the modeled polynya.

3.3.2.2 Sensitivity to Model Structure

A second set of sensitivity experiments was designed to assess the relative

importance of different aspects of the coupled ice-ocean-atmosphere system to the

observed kinematics of the Ross Sea polynya. Most of these experiments involved

manipulation of the model to eliminate certain model components. Like the sensitivity

analyses above, these experiments demonstrated that a number of physical attributes of

the Ross Sea, as well as some physical processes, play little or no role in the observed

kinematics of the Ross Sea polynya. For example, although the cavity beneath the RIS

has been shown to affect the properties of the continental shelf bottom waters {Jacobs et

aL, 1985], our results show the cavity has almost no impact on polynya formation.

Similarly, preventing solar radiation from penetrating the sea ice and entering the

water column had a very minor influence on model results, indicating that solar radiation

43

Page 57: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

warming the ocean through thin ice has a small effect compared to other processes such

as radiation warming the ocean through leads.

Likewise, removing tidal forcing had virtually no impact, despite tidal current velocities

near the continental shelf break that measured >70 cm s"1, consistent with measurements

made in the Ross Sea [Padman et al, 2003]. Some modeling studies in other Antarctic

regions have found that tidal currents significantly alter the evolution of sea ice (for

example near the continental shelf break in the Weddell Sea [Koentopp et al., 2005]).

However, the impact of tidal currents on sea ice is still under debate, and many modeling

studies have successfully simulated sea ice kinematics at a regional scale without including

tides [Ohshima andNihashi, 2005; Wu et al., 1997; Fichefet and Goosse, 1999].

Polynya development in runs where snow thickness was set to zero, either from

June through March (Figure 3.6j) or from December through March (Figure 3.6k), were

very similar to the control run, although in the absence of snow, the polynya was slighdy

larger and ice concentrations in the Eastern and Western regions were 17-50% lower than

in the control run. This indicated that the polynya was not highly sensitive to the amount

of snow cover on the ice on seasonal timescales. Finally, using daily radiation forcing

yielded model results that were almost identical to those obtained using hourly forcing.

3.3.2.3 Sensitivity to Ocean Heat

As anticipated, the sensitivity analyses revealed that the dynamics of Ross Sea

polynya formation are most sensitive to entrainment of heat from below (thermal polynya

forcing) and advection due to winds (mechanical polynya forcing). When heat

entrainment was shut off in June, the Ross Sea polynya still formed (Figure 3.6b), but it

was generally smaller and more heavily ice covered than in the control run (Figure 3.6a).

44

Page 58: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

1 January a) Control Run

1 February 1 March

b) No heat entrainment from June - March

c) No heat entrainment from December - March

d) No heat entrainment for the entire 10 year spin-up period

e) Ice not moving from June - March

f) No heat entrainment and Ice not moving from June - March

180° 140°W 140°E Longitude

180° 140"W 140°E Longitude

180" 14u°W Longitude

Figure 3.6 Austral summer sea-ice concentrations of the 10th year of the model run from (a) model control run (in which the model domain excludes the cavity under the RIS) (b) no heat entrainment June - March (c) no heat entrainment December - March (d) no heat entrainment for the entire 10 year spin-up period (e) ice not moving June - March (f) no heat entrainment and ice not moving June - March

45

Page 59: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

1 January a) Control Run

1 February 1 March

g) Open water albedo feedback effect from June - March 60°S

66°S

72°S

78°S

60°S

66°S

72°S

78°S

h) No wind June - March

i) No wind December - March

180° 140°W 140°E Longitude

20 . 40

180° 140°W 140°E Longitude

60 80

180° 140°W Longitude

[%] 100

Figure 3.6 (cont) Austral summer sea-ice concentrations of the 10th year of the model run from (a) model control run (in which the model domain excludes the cavity under the RIS) (g) open water albedo feedback effect from June - March (h) no wind from June - March (i) no wind from December - March Q no snow from June - March (k) no snow from December - March

46

Page 60: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Sea ice concentrations in the Polynya and the Western regions between 1 January and 1

March under these conditions were 49% and 32% greater, respectively, than in the

control run (Table 3.3). Turning off heat entrainment in December, just prior to the

expansion of the Ross Sea polynya, had little effect on sea ice within this region (Figure

3.6c), with mean sea ice concentrations increasing by 5% above the control run (Table

3.3). This seasonal difference in the impact of heat entrainment is to be expected because

warm MCDW flowing onto the Ross Sea continental shelf enters at depth, and is thought

to move into the surface only during deep winter mixing. During spring and summer,

increased surface ocean stratification in the region, despite being relatively weak, prevents

the warm water from mixing into the surface during that time of year \Arrigo etal., 1998].

Additionally, a run was conducted in which heat entrainment was suppressed for

the entire model spin-up of 10 years to look at the longer term effects of ocean heat

entrainment. This run shows that there is even less open water in the polynya region for

this run than there was in any of the runs in which heat entrainment was suppressed for a

shorter period of time. This is to be expected as the effects of heat entrainment can

persist for longer than annual time scales.

In general, there was much more ice present in the entire domain for the 1-year

and 10-year no-heat entrainment runs (Figure 3.6b and 3.6d). An unexpected result for

the run in which heat entrainment was shut off from June through March (Figure 3.6b)

was that mean ice concentration in the Eastern Ross Sea was 24% lower than in the

control run (0.70 versus 0.92). This is counterintuitive because eliminating heat

entrainment into the surface would be expected to result in cooler surface waters, lower

sea ice melt, and higher ice concentrations than in the control run, as was the case in the

47

Page 61: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Western Ross Sea and Polynya regions. It must be noted, however, that eliminating heat

entrainment introduced an unstable vertical density distribution into the model by

preventing warm, less dense water from entering the surface layer, an unrealistic

manifestation of this model experiment. This altered the modeled ocean circulation,

leading to a modest redistribution of sea ice in these model runs relative to the control

run. The important conclusion to be derived from these runs is that shutting off heat

entrainment resulted in an overall increase in ice concentration when averaged over the

entire model domain, despite some local decreases due to ice redistribution.

Besides being affected by ocean heat from below, open water albedo feedback

can play a significant role in the expansion of certain summer polynyas [Hunke andAckley,

2001]. The feedback occurs due to ocean water having a much lower albedo than ice,

thus, the increase in the absorption of solar energy increases the rate of polynya

expansion once there is open water. We conducted an experiment in which the albedo of

the ocean was set to 1, thereby reflecting all incoming solar radiation away from the ocean

surface. This modification had litde effect on the expansion of the modeled polynya

(Figure 3.6g), suggesting that open water albedo feedback is not an important process for

the formation of the Ross Sea summer polynya.

3.3.2.4 Sensitivity to Ice Advection

The importance of mechanical processes to the formation of the polynya was

confirmed by conducting a sensitivity run in which sea ice was prevented from advecting,

but with winds still present (Figure 3.6e). This allowed the winds to influence surface

ocean circulation, but without moving the ice. Consequently, newly formed ice would

not be able to advect downwind, sea ice concentrations would be higher in areas where

48

Page 62: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

ice initially forms, and ice concentrations would be lower in areas where ice advection is

an important source of sea ice. When sea ice was prevented from adverting, large

concentrations of sea ice formed in the Polynya (0.76) and Eastern regions (0.94) but not

in the Western region of the Ross Sea (0.0). Thus, the sea ice that is present in the

Western region in the control run must be there as a result of the advection of sea ice

from further to the south and east. More importantly, because the Ross Sea polynya fails

to form in this simulation, this polynya must form primarily through the advection, rather

than the melting, of sea ice.

3.3.2.5 Sensitivity to Winds

In contrast to the above results where the elimination of heat entrainment into

surface waters reduced the size of the Ross Sea polynya, shutting off the winds had a

much more dramatic effect, resulting in the total elimination of the Ross Sea polynya.

When winds were shut off from June through March (Figure 3.6h), no model grid cells in

the Polynya region became ice-free during the summer, the time when the Ross Sea

polynya is normally at its maximum size, and the mean sea ice concentration increased

from 0.49 in the control run to 0.96 (Table 3.3). Even when the winds were not shut off

until December, the polynya barely opened (Figure 3.6i) and ice concentrations in the

polynya region were 61% greater than in the control run. These results clearly

demonstrate that the mechanical movement of ice by wind is critical to the formation of

the Ross Sea summer polynya. In all cases where thermal processes that supply heat to

the surface ocean where eliminated, the polynya was smaller than in the control run, but it

still managed to form. Thus, while thermal processes may alter polynya size, shape, and

49

Page 63: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

timing of formation, it is advection of sea ice via the surface wind stress that is the

primary factor determining the dynamics of the Ross Sea summer polynya.

There are additional effects from suppressing wind forcing, aside from direct

effects on ice transport, which, while second order, may affect polynya dynamics. For

example, turning off the winds will reduce ocean-atmosphere heat flux because the rate

of this exchange is dependent on wind forcing. This will lead to an ocean that is, in

general, warmer than the control run in the winter and cooler in the summer,

contributing to the lack of polynya formation in this run. Additionally, surface layer

Ekman transport and wind-driven upwelling will be shut down when the wind is turned

off. However, since the movement of warm water onto the Ross Sea continental shelf is

generally driven by geostrophic flow, as opposed to wind-driven upwelling, shutting off

winds over short time scales will not likely have a direct impact on the movement of this

water mass onto the continental shelf. Although they are driven by the wind as well,

geostrophic currents operate on much longer time and space scales, and shutting off the

winds in a model run would not have as immediate an effect on geostrophic flow. Also,

shutting off the winds will not likely effect density driven vertical mixing over the

continental shelf because of the heavy ice cover present in the autumn and winter when

rates of vertical mixing are most intense.

Another run was conducted in which both heat entrainment was turned off and

ice was prevented from advecting (Figure 3.6f). Not surprisingly, this results in the region

being almost completely covered with ice during January through March and a complete

lack of polynya formation. The ice concentrations for this run are very similar to the run

in which winds were shut off between June and March (Figure 3.6h). The fact that the

50

Page 64: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

ice concentrations are greater in this run than in the runs in which either ice is prevented

from advecting or heat entrainment is suppressed further demonstrate that both thermal

and mechanical processes are important for polynya formation.

3.4 Discussion

The Ross Sea summer polynya has been modeled in the past using a 1-D

(horizontal) model of the Ross Sea [Brommch etal., 2001] as well as a global 3° resolution

three-dimensional model \Fichefetand Goosse, 1999]. The 1-D model oiBrotnmch et al.

[2001] was forced with wind and atmospheric temperature data from the Ferrell

Automatic Weather Station (AWS). This model captured some of the interannual

variability in the size of the polynya, as determined from SSM/I data. However, there are

limitations associated with using single-point forcing data and modeling die ice in one

dimension. Although their model was able to simulate polynya formation and closure

using wind forcing only, there was a temporal offset of die polynya closure, perhaps due

to the lack of ocean heat storage in their model.

The 3-D model by Fichefet and Goosse [1999] showed that the Ross Sea polynya

could not form in the absence of ice dynamics, but did not conduct a run showing the

sensitivity to entrained heat. In addition, perhaps due to the use of monthly, rather than

daily wind forcing, their modeled polynya formed one month too late, expanded too far

westward, and too much ice remained in the eastern Ross Sea in the summer. Our model

control run with daily wind forcing was able to better capture the timing of polynya

formation, as well as its spatial extent, despite a similar excess of sea ice in the eastern

Ross Sea. Neverdieless, Fichefet and Goosse [1999] were able to show that both wind-

induced divergence of the ice pack and the influence of MCDW were crucial for polynya

51

Page 65: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

formation. This is consistent with our findings that the Ross Sea polynya still formed in

the absence of entrained relatively warm MCDW, but that without it, the extent of the

modeled polynya was not as expansive as it is observed to be in satellite imagery.

In our modeling study, we investigated the relative contribution of thermal and

mechanical forcing mechanisms to the formation and summer expansion of the Ross Sea

polynya. Mechanically-driven polynyas are maintained either through wind-driven or

current-driven export of ice. Our results show that it is the transport of ice that plays the

dominant role in driving the formation of the Ross Sea polynya, with lesser contribution

from the entrainment of heat from deeper layers into the mixed layer. During the run in

which both the entrainment of oceanic heat was suppressed and the ice was not allowed

to be transported, the polynya failed to form. Likewise, when winds were shut off from

June through March, the Ross Sea polynya failed to form. However, when only oceanic

heat entrainment was shut off over the same time period, the polynya still formed,

although sea-ice concentrations were higher than in the control run. Even shutting off

winds as late as December, when the Ross Sea polynya has already begun to expand,

resulted in a large increase in sea-ice concentrations within the polynya region while

shutting off heat entrainment at this time had almost no impact (Table 3.3).

Results presented here corroborate earlier findings that both the transport of ice

due to winds and the entrainment of oceanic heat are essential to the development of the

Ross Sea Polynya in the spring and summer. Past studies of the Terra Nova Bay polynya,

which is located in the southwestern Ross Sea adjacent to the Ross Sea polynya, have

shown that katabatic winds are the dominant factor controlling the dynamics of that

polynya. For example, Bromwich et al. [1998] found that approximately 60% of winter

52

Page 66: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

polynya events were linked to katabatic surges (which diminish in January) blowing older

sea ice offshore [Bromwicb and Kurt% 1984; Van Woert et al, 2001; Brommch et al, 2001]

with a glacial ice tongue blocking ice coming from the south [Massom et al, 1998; Massom

et al, 2001]. However, the katabatic winds associated with the Terra Nova Bay polynya

average close to 16 m s"1 in the autumn and winter, much higher than those of the larger

Ross Sea polynya, which average only 6 m s"1 with little seasonal variability \Arrigo et al.,

1998]. In addition, katabatic winds near the Ross Sea polynya are not as strong or as

frequent in summer months \Arrigo et al, 1998], indicating that while they may influence

the timing of early polynya formation, they do not play a major role in the maintenance

and expansion of the Ross Sea summer polynya. Our model does not explicitly include

katabatic winds and yet it is able to simulate the formation of the summer polynya. This

suggests that synoptic-scale wind forcing, as opposed to katabatic winds, are the

dominant factor in the formation and maintenance of the large Ross Sea summer

polynya, as noted earlier by Zwallj et al [1985].

The thermally-driven component of Ross Sea polynya formation stems from the

movement of warm waters into the surface layer melting existing sea ice or preventing its

formation. Although the thermally-driven component of most Antarctic polynyas is

relatively small \Arrigo and Van Dijken, 2003], there may be an oceanic heat component to

polynyas through mechanisms such as the offshore upwelling of warm water due to

katabatic winds [Davis andMcNider, 1997] or tidal forcing such as in the Okhotsk Sea

[Polyakov and Martin, 2000]. Jacobs and Comiso [1989] suggested that heat input from the

ocean was at least partially responsible for maintaining open water in the Ross Sea, as

they felt that winds alone could not explain the polynya advancing at a rate of 40 km d '

53

Page 67: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

0 25 50 75 100 Percent of time ice covered

Figure 3.7 Percent of time in November and December during which ice concentrations were greater than 10% (satellite data from 1997 - 2001). The position of the 1000m isobath is shown by the black line. \Arrigo and Van Dijken, 2004]

in December when winds at a nearby weather station averaged 3 m s"1 and air

temperatures were less than -6°C.

Jacobs and Comiso [1989] also noticed that ice cover over the continental shelf was

persistentiy lower than above the adjacent deep ocean. This finding was later confirmed

hjA.rtigo and Van Dijken [2004], who demonstrated a remarkable correspondence

between waters with a low number of days with ice cover and the shallow bathymetry

associated with the continental shelf (Figure 3.7) \Arrigo and Van Dijken, 2004]. Although

the reduced ice concentrations in these shallow waters have been attributed to the flow of

warm MCDW onto the continental shelf, whether this warm water melts existing sea ice

or simply inhibits the formation of new sea ice is unclear. It is unlikely, however, that

MCDW is entrained into the surface during the summer to cause ice melt at that time of

54

Page 68: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

year. For example, hydrographic data for the Ross Sea show a minor amount of

meltwater in surface waters in the Ross Sea [Smith and Gordon, 1997], which remain weakly

stratified throughout the summer \Arrigo et al, 1998; Arrigo et al, 2000]. Furthermore,

Jacobs and Comiso [1989] concluded that while MCDW enters the shelf year-round, it only

retards ice growth when entrained into the vertical circulation pattern during the winter.

The lack of evidence for ice melt for surface waters in the Ross Sea in the summer is

further indication beyond our model results that the open water albedo feedback effect

does not play a large role in the formation of the Ross Sea summer polynya. Thus, this

polynya is fundamentally different from other polynyas in which this feedback effect has

a larger impact. This is likely due to the dominating influence of winds in the formation

of the Ross Sea summer polynya, which advect sea ice out of the southern Ross Sea

before it has a chance to melt appreciably.

Our conclusion that subsurface oceanic heat is entrained only during winter

mixing and not during the summer is consistent with the lack of ice melt and ocean

stratification signatures observed in the Ross Sea in the summer. Our modeled vertical

profiles of temperature (Figure 3.8) show warm MCDW at depth both on and off the

shelf, with a cold water mass located above the MCDW over the continental shelf.

Thermal stratification intensifies during the summer, as sea surface temperatures warm

from December to March when the polynya is open. As a result, warm MCDW on the

shelf remains clearly separated from warm surface waters during spring and summer

(Figure 3.8). It is only during winter when convective overturn vertically mixes the water

column over the continental shelf that the heat from MCDW is entrained into surface

waters. These results suggest that the movement of warm MCDW onto the shelf, and

55

Page 69: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Mar

ch

Febr

uary

Ja

nuar

y D

ecem

ber

v>

tn

•̂h o rt>

tern X)

rt>

:ratu n O

< H B-

n 3 o del a.

o 3 5

? o/5' e ft

l» bo

g o a- &

. a- ®

< B

. o e.

n 3 TS 8 0 H l-t

0 &

a>

CO

p r+

i_

i

-0

Ui o w

a- ^

NJ M

O)

aj

Co

o>

o

CO

~-i

00

CO

-4

lo

CO

o

Page 70: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

the entrainment of its heat into surface waters via vertical mixing in the winter, reduces

the rate of winter ice growth, rendering the sea ice more easily advected to the north

\Dinniman et al, 2003; Gordon et al, 2000; Arrigo and Van Dijken, 2004; Jacobs and Comiso,

1989].

While the model used here has proven to be a useful tool for understanding the

dynamics of the Ross Sea summer polynya, there are some discrepancies between

modeled sea ice and observations, with respect to both ice concentrations and

thicknesses, that can be addressed in future studies. The most likely causes of these

discrepancies include modeled deep ocean temperatures that are slightly higher than

reality is some locations and lower in others, the absence of modeled katabatic winds, and

the lack of snow redistribution in the model. In general, modeled ice was thicker than

observed, and this may be attributable to the model producing too little oceanic heat.

However, because there is not a robust observational data set for deep ocean

temperatures throughout the Ross Sea, it is not yet possible to assess the performance of

the model in this regard (although the model does agree with what little deep ocean

temperature data are available). On the other hand, the discrepancies could merely be

attributable to interannual variability in snow and ice distributions that is not captured by

the model. Low sea ice years would cause melting of multi-year ice and lead to thinner

ice the following year. In addition, we have shown that altering snow accumulation on

the sea ice has a minor impact on ice concentrations in this study region. However, the

model does not presently include redistributions of snow due to wind, which may be

important in the overall snow mass budget [Dery and Tremblay, 2004] and could affect sea

ice concentrations to a larger extent than has been suggested here. Finally, as discussed

57

Page 71: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

previously, the model does not include extreme wind speeds associated with katabatic

wind surges, which may be important in the timing of the polynya formation. These

peaks in wind speed could be particularly important since wind stress is proportional to

the square of the wind speed. However, some studies have looked at the factors

contributing to the timing of the Ross Sea polynya and have found that autumn and

winter air temperatures have a greater correlation to polynya formation than the

magnitude of the wind stress \Arrigo eta/., 1998], suggesting that katabatic winds are not

critical to polynya development.

3.5 Conclusions

Our model captured the formation of the Ross Sea polynya with respect to both

its spatial extent and temporal evolution and corroborated earlier findings that both

thermal and mechanical processes play a role in the formation of this polynya. We

showed that ice advection, due mostly to wind forcing as well as ocean currents, is the

most important driver of polynya formation. The entrainment of heat from warm

MCDW in winter also plays a substantial role in polynya development, with warm water

likely moving onto the continental shelf year-round and retarding ice growth in the

winter. This allows the ice to be advected away more easily when stronger winds come in

the spring and summer.

58

Page 72: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 4 Ross Ice Shelf Cavity Circulation, Residence Time, and Melting: Results from a Model of Oceanic Chlorofluorocarbons

Abstract

Despite their harmful effects in the upper atmosphere, anthropogenic

chlorofluorocarbons dissolved in seawater are extremely useful for studying ocean circulation and

ventilation, particularly in remote locations. Because they behave as a passive tracer in seawater,

and their atmospheric concentrations are well-mixed, well-known, and have changed over time,

they are ideal for gaining insight into the oceanographic characteristics of the isolated cavities

found under Antarctic ice shelves, where direct observations are difficult to obtain. Here we

present results from a modeling study of air-sea chlorofluorocarbon exchange and ocean

circulation in the Ross Sea, Antarctica. We compare our model estimates of oceanic CFC-12

concentrations along an ice shelf edge transect to field data collected during three cruises

spanning sixteen years. Our model produces chlorofluorocarbon concentrations that are quite

similar to those measured in the field, both in magnitude and distribution, showing high values

near the surface, decreasing with depth, and increasing over time. After validating modeled

circulation and air-sea gas exchange through comparison of modeled temperature, salinity, and

chlorofluorocarbons with field data, we estimate that the residence time of water in the Ross Ice

Shelf cavity is approximately 2.2 yr and that basal melt rates for the ice shelf average 10 cm yr1.

The existence of a mode in the model during which warm modified circumpolar deep water

directly enters the ice shelf cavity suggests that this system may be particularly sensitive to small

perturbations that could influence circulation patterns.

59

Page 73: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

4.1 Introduction

As we proceed into a world of unprecedented climatic changes, it is critical to

understand the mass-balance of Antarctic ice shelves, as these bodies of ice are the

keystone to the sea-level changes that are likely to be caused by the melting of Antarctic

ice sheets [IPCC AR 4 2007]. Because the ice shelves themselves are floating, they are

capable of only minor direct contributions to sea level variations [Jenkins and Holland,

2007]. However, these grounded shelves slow the movement of ice from the continental

interior into the ocean. This continental ice could cause sea level to rise by as much as 60

m if the entire Antarctic Ice Sheet disintegrated [IPCC AR 3 2001]. The mass balance of

the ice shelves is affected positively by the flow from this continental ice sheet, by

precipitation, and by basal freezing along the bottom of the ice shelf. Mass is lost most

substantially by basal melting \Hellmer, 2004; Jacobs et al, 1992] as well as by surface

ablation and iceberg calving. Presently, these ice shelf systems are not well understood

due to our limited ability to directly observe the base of the ice shelves as well as the

circulation patterns and temperatures of seawater within the cavities below.

Anthropogenic chlorofluorocarbons (CFCs, Table 4.1) are a unique and valuable

ocean tracer, despite their detrimental role in stratospheric ozone depletion. Their

concentrations in the troposphere are well-mixed and the amounts reflect their increase

from non-existent to substantial levels during the last 70 years (Figure 4.1)

Table 4.1 Chi

CFC-11 CFC-12 CFC-113

oroflourocarbo formula CC13F CC12F2

C12FC-CC1F2

n descriptions Full name trichlorofluoromethane dichlorodifluoromethane trichlorotrifluorethane

other names f reon- l l ,F- l l ,R- l l freon-12, F-12, R-12 freon-113

60

Page 74: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

[Walker et al, 2000]. Atmospheric CFCs dissolve in seawater, and their gas-exchange

coefficients have been experimentally determined [Bu and Warner, 1995; Warner and Weiss,

1985]. Starting in the early 1990s, CFCs have been used to evaluate ocean circulation in

the Ross Sea \Trumbore et al, 1991]. They are especially useful in studying sub-ice shelf

cavities that are isolated from the atmosphere and consequently have no additional source

of CFCs. Thus, vertical profiles of CFC concentrations measured along the mouth of the

cavities can be used to determine where water is entering and exiting and can be used as

proxies for cavity residence times. Since making measurements beneath the Ross Ice

Shelf (RIS) is extremely difficult, in the present study we have used observed changes in

CFC-12 distributions in conjunction with numerical modeling to understand the

S I 31 o

u LL. U

600

500

400

300

200 '

100

CFC-12

— CFC-11

~ - CFC-113

0 1940 1950 1960 1970 1980 1990 2000

Year

Figure 4.1 Atmospheric concentrations in the Southern Hemisphere of CFC-12, CFC-11, and CFC-113 from 1940 to 2000. [Walker et al. 2000]

61

Page 75: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

circulation, residence times, and ultimately the melting rates within the RIS cavity. A

coupled ice-ocean-atmosphere model was used to estimate CFC gas transfer between the

atmosphere and the ocean as well as the advection-diffusion of CFC concentrations in

seawater. A robust dataset of CFCs (Figure 4.2) measured along the edge of the ice shelf

between 1984 and 2000 [Smethie and Jacobs, 2005] was used to validate modeled air-sea

exchange and ocean circulation on the continental shelf and within the cavity. The model

was then used to estimate the difficult-to-measure basal melt rate of the RIS as well as the

residence time of water within the RIS cavity.

Figure 4.2 (a) Map of the Antarctic coastline and the 1000 m isobath. The Ross Sea model domain perimeter is also shown, (b) CFC sampling locations along a transect following the RIS edge [Smethie and Jacobs 2005] during Jan-Feb 1984, Feb 1994, and Feb-Mar 2000.

62

Page 76: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

4.2 Modeling Methods

We modeled the Ross Sea region using the Polar Ocean Land Atmosphere and

Ice Regional (POLAIR) modeling system [Reddy et al., 2007; Holland and Jenkins, 2001;

Holland et al, 2003], which includes components representing ocean circulation, land

margins (including coastlines), ice thermodynamics-dynamics (including glacial and sea

ice), and a simple atmosphere.

4.2.1 Ocean Component

The Miami Isopycnic Coordinate Ocean Model (MICOM) [Bleck and Smith, 1990],

an isopycnic, primitive equation code, was modified for use as the ocean circulation

component of the model. Parameterizations allow for weak diapycnic mixing between

vertically adjacent isopycnic model layers as well as strong lateral mixing along isopycnic

layers. To a large extent, these parameterizations mimic the observed mixing behavior in

the real ocean [Holland and Jenkins, 2001]. The Circum-Antarctic Tidal Simulation (CATS)

model [Robertson et al, 1998] was used to simulate local non-equilibrium tidal forcing. The

modeled ocean floor, the surface of the sub-ice shelf cavity, and the initial density

structure of the ocean were defined by datasets as described by Holland et al. [2003]. A

sponge zone of three grid cells is used to restore temperature and salinity to a prescribed

pattern based on climatology (see Holland et al. [2003] for details) along the northern,

eastern, and western sidewalls of the open ocean part of the domain. The interior-most

sponge cells are restored on a 30-day timescale for baroclinic fields and the exterior-most

are restored on a 10-day timescale. For barotropic fields, timescales of 1-day and 5-days

are used for the interior and exterior sponge cells, respectively. This approach helps to

prevent artificial drift due to the model's limited spatial domain, provides reasonable

63

Page 77: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

lateral boundary conditions, and ensures that properties at the boundaries are kept close

to observed values. No restoring was used for sea surface salinity other than along the

sidewalls of the domain.

4.2.2 Sea-Ice and Ice Shelf Components

Two time varying ice thickness categories were used in each grid cell for the sea-

ice component of the model; one represents ice floes and the other represents leads

between ice floes. A cavitating-fluid rheology \Flato andHibkr, 1992] is used to treat ice-

ice interactions whereby the ice pack is resistant to convergence but not divergence. We

compute sea-ice concentration and thickness by accounting for the divergence of heat

and freshwater fluxes at the interfaces between the sea ice and the ocean and atmosphere.

Thus, in a number of ways, the sea ice is modeled as an "isopycnic" fluid layer, similar to

the ocean, but with different stress properties [Holland and Jenkins, 2001]. A complete

discussion of ice thermodynamics can be found in Holland et al. [1993] and Holland

[1998]. Sea ice concentration is computed for each grid cell as a fraction (0-1) of total

grid cell area. Initial polynya formation and its subsequent expansion in the summer in

the Ross Sea is simulated well using the POLAIR modeling system [Reddy et al, 2007].

Precipitation over sea ice increases snow thickness based on the density of snow

(300 kg m'3) with precipitation derived from a global climatology [Legates and Willmott,

1990]. Snow thickness decreases when snow sublimates or melts, or when snow is

transformed into sea ice. This latter, snow-ice production occurs through processes such

as flooding when there is sufficient snow cover to submerge the sea ice below freeboard.

Thermodynamic interactions at the interface between the ice shelf base and ocean

surface, as well at the sea ice-ocean interface, are described using a viscous-sublayer

64

Page 78: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

parameterization [Holland andJenkins, 1999]. To incorporate the arbitrary surface

topography of an ice shelf base, the ocean model was modified by Holland and Jenkins

[2001]. The ocean cavity below the RIS is modeled as described by Holland et al.[2003].

4.2.3 Atmosphere and CFC Gas Exchange

For air temperature and sea-level pressure, NCEP reanalysis products [Kakay et

a/., 1996] are used as atmospheric forcing. Using the horizontal gradient of daily sea-level

pressure, the model computes daily geostrophic winds. In the absence of a robust dataset

for relative humidity over the ocean, a value of 0.90 was used by the model. A constant

cloud cover of 0.85 was used based on satellite-derived estimates of cloud cover over

open water in the Ross Sea for the period of November through February \Arrigo and van

Dijken, 2004].

Atmospheric CFC data from Walker et al. [2000] were used to estimate the air-sea

flux of CFC-12. The flux formulation is derived from the one-dimensional form of

Pick's first law, given as follows [Uss and Slater 197'4]:

F = KAC 1

where F (in units of mass [length2 time]"1 or MfL2!]"1) is the flux of gas through the layer,

K [LT1] is the exchange constant, and AC [ML3] is the concentration difference across

the layer thickness. Given that the gas exchange obeys Henry's Law:

P = HC 2

where P [patm] is the partial pressure of the solute, C [pmol kg4] is the concentration of

the solute in solution, and H Peg patm pmol"1] is Henry's law constant. Fick's law can be

reformulated, as follows:

65

Page 79: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

pair

F = KAC = Kwater(-^-C^) = Kwater(aPc%-Cwc^) 3

where Kmter [m s"1] is the exchange constant for seawater (the term is also referred to as

the gas transfer coefficient or piston velocity), P££c [patm] is the partial pressure of CFC

in dry air at a pressure of one atm, C ^ T [pmol kg1] is the CFC concentration in

seawater, a [pmol (kg patm)"1] is the reciprocal of H which, in this case, represents the

CFC solubility for water-vapor saturated air (also referred to as the partial pressure

equilibrium constant). Also note that P£pC is equivalent to the dry air mixing ratio

multiplied by 1012. Thus, the oceanic CFC concentration at the next modeled time step,

i+1, is computed as:

csra+i)=csr(o+^wlrtr[a^(o-csr(o]^ 4 An

where A/ [s] is the calculation timestep and Ah [m] is the thickness of the ocean mixed

layer. The piston velocity, K„ter, is calculated as follows [Wanmnkhof, 1992]: & water ~ Q 1.6ES[™»/mhr} ' V4 + V ) " JSc/660 5

where a [cm s~2(hr m2)"1] is a constant of 0.31 as given by Wanninkhoj'[1992] for steady

winds, tt [m2 s"2] is the wind speed squared and v (m2 s"2) is the wind speed variance. Sc

(dimensionless) is the Schmidt number which is temperature dependent and given as:

Sc = ax + a2Tc + a3Tc2 + a4Tc

3 C

where X, [°C] is the temperature and at ate the diffusion coefficients [Haine and VJchards,

1995; Zheng etal., 1998], and differ for each CFC gas (see Table 4.2). CFC solubility, a, is

calculated using the formulation given by Warner and Weiss [1985]:

66

Page 80: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

a = exp(4 + A, + 4 log TKim + AJK/m2 + s(d(B1 + B2TKim

+ "3* Kim )) 7

JK/100

where A^ and B; are coefficients summarized in Table 4.2 [Bu and Warner, 1995; Warner and

Weiss, 1985], TK/W0 is the ocean temperature in Kelvin divided by 100, and salis salinity.

We assumed that there was no atmosphere-ocean CFC exchange through sea ice,

leading to a CFC gas-exchange flux that is proportional to the fraction of open water in

each grid cell.

Sch

mid

t N

umbe

r (S

c)

Gas

Sol

ubili

ty

for CFC-11: a, = 3501.8 a2=-210.31 a3= 6.1851 a4= -0.07513

A, = -232.0411 A2= 322.5546 A3= 120.4956 A4= -1.39165 B,= -0.146531 B2= 0.093621 B,= -0.0160693

for CFC-12: Jh = 3845.4 a2=-228.95 a3= 6.1908 a4= -0.067430

Aj = -220.2120 A2= 301.8695 A3= 114.8533 A4= -1.39165 B, = -0.147718 B2= 0.093175 B3= -0.0157340

fof CFC-113:

^ C C F C - 1 1 3 — 1 - 1 6 ^CCFC-11

A, =-231.902 A2 =322.915 A3=119.111 A4 =-1.3917 Bj = -0.02547 B2 = 0.00454 B3 = 0.0002708

Table 4.2 Constants for the temperature-dependent Schmidt number [Zheng 1998 (CFC-11 and CFC-12); Haine and Richards, 1995 (CFC-113)] and gas solubility [in units of pmol kg'1 patnY1, Warner and Weiss 1985 (CFC-11 and CFC-12); Bu and Warner 1995 (CFC-113)]

4.2.4 Model Grid and Runtime

The horizontal model grid is based on a Mercator projection and all grid cells are

rendered to be locally square. The horizontal grid box size in the middle of the Ross Sea

domain is approximately 2°. Moving southward in the model domain, grid cells become

smaller in proportion to the cosine of the latitude (2° zonally and 2°cos(cp) meridionally,

where cp is latitude). The model domain reached from 155°E to 135°W and from 60°S to

67

Page 81: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

84°S, with 24 vertical layers. This domain includes die Ross Sea continental shelf, die

deep ocean, and die cavity beneatii die RIS (Figure 4.2).

The model was run over the simulated time period 1950-2000. Although there

were detectable levels of CFC in the water column prior to 1950, during the first five

years of model spin up, CFCs remained low, less dian 3% of tiieir maximum values.

(Sensitivity analyses show that starting the model runs with detectable levels at diis point

in time does not have a significant effect on modeled CFC concentrations by 1984.)

4.2.5 Residence Time

We estimated die residence time of water witiiin the RIS cavity in the model using

an artificial dye tracer approach. Initial dye concentrations were set to 1.0 in all grid cells

beneath die RIS and set to zero in all grid cells outside the RIS. The dye was tiien

allowed to advect and diffuse by the ocean model as a passive tracer, and die amount of

time it took for the mean dye concentration witiiin the RIS cavity to be reduced to 0.5

was taken to be the half-life of dye witiiin the cavity. The reduction of dye in die cavity

resembles exponential decay, so the residence time (x) can be computed as:

where, tVl is die half-life of dye in RIS the cavity \Killops andKillops, 2005].

4.3 Results and Discussion

4.3.1 Modeled CFC Distributions

In accord witii all three years of field data from the Ross Sea [Smethie and Jacobs,

2005], the model produces CFC concentrations (Figure 4.3) that are highest in the surface

68

Page 82: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

waters (-2.0 pmol kg1 in 1984, -3.0 pmol kg1 in 1994, and -3.5 pmol kg1 in 2000) and

decrease significandy towards the bottom (-1.0 pmol kg"1 in 1984, —2.0 pmol kg4 in

1994, and -2.5 pmol kg"1 in 2000). However, the model, which creates a thick surface

layer of high CFC concentrations, does not capture the strong surface stratification that is

observed. The model also confirms the noteworthy seasonal signature in CFC

concentration as the CFCs move from the atmosphere into the mixed layer (Figure 4.4),

with the highest flux and greatest concentrations in the surface waters in the late austral

summer. The high winds and absence of sea ice facilitates enhanced CFC exchange with

the atmosphere despite the fact that warmer water temperatures decrease gas solubility.

Deep mixing in the winter distributes the CFCs throughout much of the water column

and dilutes the high surface summertime concentrations.

Evident in the field dataset for CFC-12 in 1984, 1994, and 2000 is a region of

low-CFC water at (200-800 m) depth centered between 178°W and 175°W, as well as a

slightly less pronounced low-CFC region that is shallower (100-500 m) and centered

around 170°E (Figure 4.3). These features have been suggested to reflect a plume of low-

CFC water diat exits the RIS cavity after prolonged isolation from atmospheric CFC

sources [Smethie and Jacobs, 2005]. Similar plumes of low CFC water can be seen in the

model results (Figures 4.3), although they are formed at slighdy different depdis and

locations. Both plumes are formed deeper in the water column in the model, with die

westward plume centered near 165°W and the eastward plume approximately 100 m

above the ocean floor and centered around 178°E. The horizontal distributions of

modeled CFCs at 450 m depth (Figure 4.3c) show a mass of low-CFC water to the north

of die cavity mouth that appears to intrude into the cavity.

69

Page 83: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

(a) o 1984 (Early February) 1994 (Mid February) 2000 (Early March)

140E 160E 180 160W 140W 140E 160E 180 160W MOW 140E 160E 180 160W 140W

longitude

Figure 4.3 CFC-12 concentrations [pmol kg-1] from (a) field studies along a transect following the Ross Ice Shelf edge from Jan-Feb 1984; Feb 1994; Feb-Mar 2000 [Smethie and Jacobs 2005], (b) model estimates along 77.3°S for equivalent model dates, (c) model estimates at 450 m surrounding the mouth of the cavity, (d) model estimates at 450 m depth for the entire model domain, and (e) model estimates for the upper mixed layer for the entire model domain.

70

Page 84: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Figure 4.4 Model estimates of CFC-12 concentration [pmol kg4] at 180°, 77.3°S from 1980 through 2000.

Modeled mixed layer CFC concentrations (Figure 4.3e) clearly increase over time between

1984 and 2000. Surface concentrations correspond in part to sea-ice distributions, with

lower concentrations located within the Ross Sea polynya than the surrounding ice

covered waters, particularly in 1984 and 2000. The lower values occur when deeper

convective mixing on the shelf dilutes high surface concentrations. These surface values

reflect a near equilibrium state with the atmosphere, which is a function of water

temperature and salinity, with concentrations also being affected by lack of gas-exchange

in the presence of sea ice as well as dilution of CFCs when the mixed layer thickness is

large.

71

Page 85: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

150E 160E 170E 180 WOW 160W 150W HOW 130W

Longitude -^

150E 160E 170E 180 170W 160W 150W HOW 130W

Longitude -£>

1

0.7

0.4

0.1

-0.2

-0.5

-0.8

-1.1

-1.4

-1.7

- 2

Figure 4.5 Modeled temperature and current velocity (averaged over the final year of the model run) for (a) the surface layer and (b) 450 m depth.

4.3.2 Modeled Circulation, Salinity, and Temperature

Mean ocean currents were computed for the entire model domain, including the

cavity beneath the RIS (Figure 4.5). Velocities in the surface layer and at a depth of 450

m were averaged over the final year of the model run. There is considerable seasonal and

interannual variability in the velocity fields. Velocities on the open continental shelf are

in the range of 2.0 to 8.0 cm s"1, with much lower velocities in the RIS cavity of between

0.2 and 1.0 cm s4. The velocity vectors clearly show the clockwise Ross Sea gyre located

72

Page 86: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

to the northeast of the continental shelf with velocities of approximately 2-3 cm s"1.

Waters run westward along the edge of the RIS at velocities of <1 cm s"1 and then

northward along the edge of the continent until it flows off the continental shelf.

Eventually, these waters flow westward along the coast as the east wind drift. At a depth

of 450 m, a mass of warm Circumpolar Deep Water (CDW) originating north of the

continental shelf can be seen to move southward onto the eastern side of the shelf,

transforming into Modified Circumpolar Deep Water (MCDW) by mixing with cooler

ambient waters.

Within the RIS cavity, velocities are much lower than they are on the shelf,

reflecting the absence of direct wind forcing on waters within the cavity. Waters generally

move in a clockwise direction, from the east to the northwest, both in the surface layer

and at 450 m depth. At both depths, there is stronger flow, just under 1 cm sA, in the

eastern side of the cavity as well as in the southern most reaches. In the surface layer, this

flow moves to the northwest more slowly (approximately 0.25 to 0.5 cm s"1) than in the

east, exiting the cavity in the west. At 450 m, this flow moves to the north in the center

of the cavity at about 0.5 cm s4. Along the west side of the cavity, there is almost no

movement of the water (flow < 0.1 cm sA) at 450 m. At this depth, flow exits the cavity

in the center of the RIS transect.

The persistent salinity pattern along the RIS transect both in situ and in the model

(Figure 4.6) is marked by a mound of salty deep water in the western side of the study

region, with the halocline deepening rapidly between 170°E and 170°W. This mound of

salty water is maintained in geostrophic balance by the strong northward flow along the

west side of the model domain. Maximum salinities in the deep western section of the

73

Page 87: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

1984 (Early February) 1994 (Mid February) 2000 (Early March)

luuu 170E 180 170W 160W 170E 180 170W 160W 170E 180 170W 160W

Figure 4.6 Salinity from (a) field studies (along a transect following the ice shelf edge Jan-Feb 1984; Feb 1994; Feb-Mar 2000) and (b) model estimates (along a transect of 77.3°S for equivalent model dates)

transect are somewhat higher in situ than they are in the model (34.8 and 34.6,

respectively). Like the freshening of deep waters observed in the field data over the two

decades [Smethie and Jacobs, 2005], there is a slight freshening in the model, with values

above 34.65 produced in 1984, but not afterwards.

In situ temperatures are in the range of -2°C to -1.4°C along the RIS transect for

the years sampled (Figure 4.7a), with pockets of higher temperature (-1.4°C to 0°C)

observed in surface waters. Surface temperatures were lower along the 2000 transect,

which may be attributed to sampling taking place in late rather than mid-summer.

Modeled temperatures along the RIS transect (Figure 4.7b) range from -2°C to -1.4°C

over most of the region, broadly similar to observations. However, the model does not

produce the same degree of warming in the surface layers (-1 to 0°C) and exhibits areas

of relatively warm water at depth (-1.4 to -0.2°C). These plumes of warm water originate

in the north, as can be seen in the horizontal section of CFCs at 450 m depth

74

Page 88: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

1984 (Early February) 1994 (Mid February) 2000 (Early March)

170E 180 170W 160W 170E 180 170W 160W 170E 180 170W 160W • -2.0

Longitude

Figure 4.7 Temperature from (a) field studies (along a transect following the ice shelf edge Jan-Feb 1984; Feb 1994; Feb-Mar 2000) and (b) model estimates (along a transect of 77.3°S for equivalent model dates) (c) model estimates surrounding the mouth of the cavity

surrounding the cavity mouth (Fig. 7c), and represent MCDW. Although this warm deep

water was not evident in the in situ data collected in 1984 and 1994, it was observed in

the 2000 field data as well as in previous studies of Ross Sea hydrography and has been

attributed to MCDW moving over the shelf [Jacobs and Comiso, 1989; Jacobs and Giulivi,

1999; Jacobs et al, 1970; Jacobs etal, 1979; Smethie and Jacobs, 2005]. These plumes of

MCDW have been observed before in the southern Ross Sea at a depth of 250 m

centered at approximately 165°W [Jacobs et al., 1979], 175°W [Jacobs and Giulivi, 1999] and

175°W [Jacobs et al., 1970]. Furthermore, Jacobs and Comiso [Jacobs and Comiso, 1989]

observed two distinct plumes of warm water, similar to those produced in the model,

along 173°W at 250 m and at 178°E at a depth of 200 m. The model predicts two

plumes of deep warm water, one centered at 180° at a depth of 400 m and the other

centered along 165°W at 500 m.

75

Page 89: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

4.3.3 Residence Time Estimates

Over the course of the model run, the dye tracer of water in the cavity dilutes

significantly near the mouth of the cavity with much higher concentrations in the far

southern reaches of the cavity (Figure 4.8). There is a fairly smooth gradient of increasing

concentrations towards the interior of the cavity, with greater intrusion of water into the

cavity in the east. It can clearly be seen that the residence times are greatest for water at

the far south of the cavity and decrease moving northward. The dye also shows that

there is flow into the cavity along the east and out of the cavity along the west, consistent

with modeled velocities and the flow pattern that we have described above. We estimate

that the mean residence time for water in the entire cavity is approximately 2.2 yrs. This

age is consistent with previous residence time estimates for the upper layer of the RIS

cavity of <6 yrs made 400 km south of the ice shelf edge at the Ross Ice Shelf Project Ice

Hole (82°22.5'S,168°37.5'W) [Michelet al., 1979]. It is also within with the range of 0.9-

5.4 yrs calculated by Smethie and Jacobs [2005], which was based on box and stream tube

model calculations constrained by observed CFC concentrations, although our estimate is

somewhat lower than their best fit of 3.4 yr.

4.3.4 Low-CFC Waters

Examining the in situ temperatures (Fig. 6a) associated with the low-CFC waters

at the mouth of the RIS cavity (Fig. 3a), it can be seen that there is no correlation

between CFC concentration and water temperature in the observations. In 1984, the

low-CFC waters were associated with both a region of cold water marked by the -1.92°C

isotherm and a pocket of warm water (-1.1°C) centered near 174°W at 300 m depth.

76

Page 90: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Figure 4.8 Distribution of RIS cavity "dye" in the mixed layer at the half-life of the dye in the cavity [%]

During 1994, the deep eastern low-CFC waters (centered near 178°W and 500 m depth)

correspond very well with the extremely cold -1.92°C isotherm, while the shallower

western low-CFC water (centered near 174°E and 200 m depth) was co-located with

warmer temperatures of up to -1.2°C. Finally, in 2000, the low CFC region centered near

174°W corresponded with warmer waters (-1.6 to -1.2°C) while the eastern low-CFC

region was associated with relatively cold water.

In contrast, low-CFC waters in the model consistendy correspond to regions with

warm ocean temperatures. This is unexpected as the low-CFC regions along the RIS

transect are often associated with water leaving the RIS cavity [Smethie and Jacobs, 2005],

which is colder than the water on the open continental shelf since much of its heat has

been extracted as it moved within the ice shelf cavity. Therefore, model results suggest

that rather than representing waters moving northward from beneath the RIS, the low

CFC concentration in waters along the RIS edge also may represent the advection of

warm CDW southward onto the shelf, where it mixes with shelf waters to become

MCDW, and into the RIS cavity. CDW is low in CFCs due to it's isolation from the

77

Page 91: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

atmosphere [Smethie and Jacobs, 2005]; thus warm water moving onto the shelf and

becoming MCDW would be similarly low in CFCs. Because waters flowing out from the

RIS are relatively cold (Figure 4.7), the plumes of warm, low-CFC water observed along

the mouth of the cavity are entering the cavity instead of exiting. (This is confirmed by

model results showing that the warm, low-CFC plumes along the mouth of the cavity

originate from outside the cavity.) The warm salty CDW moves onto the Ross Sea

continental shelf at a depth of 400 to 500 m and flows southward along the ocean floor

[Jacobs andComiso, 1989], occasionally entering the mouth of the cavity at depth while

retaining its characteristic low-CFC concentrations, especially in the east where the

continental shelf is narrow.

Although the field data do not show a clear signature of warm MCDW entering

the cavity in 1984 and 1994, die data from 2000 show warm, low-CFC water at depth that

appears to be entering the cavity. Additionally, there are many observations showing

warm MCDW at depth on the Ross Sea Continental Shelf [Jacobs andComiso, 1989; Jacobs

andGiulivi, 1999; Jacobs eta/., 1970; Jacobs eta/., 1979]. Furthermore, warm CDW is known

to enter the Pine Island Ice Shelf cavity [Jacobs et a/., 1996] (located to the northeast of the

larger RIS cavity), and is thought to be a risk to the melting of the ice shelf. Our model

results, in conjunction with observations, suggest that there are two distinct modes of

circulation on the Ross Sea continental shelf. In the first mode, high salinity shelf water

(HSSW) is found at depth along the mouth of the cavity. Water from this source is

formed by the free2ing of seawater on the continental shelf and is very cold and dense,

consistent with what is observed in field data. This mode does not involve the flow of

78

Page 92: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

low-CFC MCDW into the RIS cavity. The second mode still involves formation of

HSSW, but is dominated by warm MCDW directly entering the cavity at depth.

4.3.5 Basal Melt Rates

Melting of the RIS is greatest near the cavity mouth (Figure 4.9), due to the heat

sources from outside the cavity, including the warm, deep, MCDW described above and

surface waters warmed by the sun and the atmosphere. The highest rates of melting

occur in September and March. In September, melting along most of the cavity mouth is

approximately 0.3 m monthA, with rates between 0.08 and 0.3 m month"1 extending into

the cavity 100 to 200 km. After September, melting rates drop until around January, and

then increase again until March. In March, there is a similar pattern of melting as in

September, with melt rates exceeding 0.35 m month4, but with fairly high degrees of

79

Page 93: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

melting extending over a smaller area. These high melting rates near the mouth of the

cavity are in contrast to estimates of basal freezing that dominate most of the interior of

the RIS, where ice accretion rates range from 0 to 0.04 m month"1 throughout the year.

Overall, basal melting for the cavity is lowest in May through July, with fairly uniform

freezing both in the interior of the cavity and near the mouth during this time. Taking

both ice shelf basal melting and freezing into account, the mean annual modeled basal

melt rate for the RIS is 0.10 m yr"1. This rate is comparable to the previous estimate of

—0.15 m yr"1 for areas of the RIS that exceed a 300 m draft [Smethie and Jacobs, 2005].

Although model estimates of high basal melting in the winter may seem

counterintuitive, higher melting rates are the result of more intense convective mixing on

the continental shelf in the winter. As salt rejection due to the refreezing sea ice

destabilizes the water column, deep warm MCDW is entrained into near surface waters.

This additional heat source contributes to greater basal melting of the ice shelf during the

otherwise cold winter. In March, the melting rate is also high, in this case due to warm

surface waters on the Ross Sea continental uncovered by sea ice and exposed to solar

insolation and increased air temperatures. These waters enter the RIS cavity in the

surface and contribute to the high melt rates at this time of year.

Despite the fact that the flow of MCDW into the RIS cavity was only observed

during the 2000 field season, its existence in the model suggests that the system may be

quite sensitive to a perturbation away from the current mode (in which HSSW is found at

depth along the mouth of the cavity) towards an alternative state. In this state, there

could be increased melting at the base of the RIS due to the MCDW entering the cavity

year round and then in cold winter months, the freezing of sea ice just outside the cavity

80

Page 94: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

causing the water column to mix more deeply, distributing some of the warm MCDW at

depth into the surface. Similarly, warm water is known to enter the Pine Island Ice Shelf

cavity and melt the ice shelf there {Jacobs etal., 1996]. It is critical to acknowledge the

potential for this warm water to move into the RIS cavity, because depending on die

severity, such events could have a destabilizing effect on the Ross Ice Shelf. Currendy, the

observed frequency and magnitude of warm MCDW entering the RIS cavity is less than

die model estimates. The CDW water mass is old, connected to the North Adantic

through thermohaline circulation, and has been at depth too long for its temperature to

be affected by recent global warming. However, it is possible that climate change will

alter circulation patterns on the Ross Sea continental shelf. One mechanism by which this

may occur is that westerlies are expected to shift due to global warming [Russe/iet al.,

2006]. It is understood from the paleorecord that a latitudinal shift in the westerlies could

substantially affect upwelling of deepwater masses in the Southern Ocean \Lamj et al.,

2007]. Such an increase in upwelling could put pressure on the oceanographic system in

die Ross Sea towards the mode exhibited in the model in which there is a larger amount

of warm MCDW entering the RIS cavity. This could cause a greater degree of melting of

the base of the RIS, which could result in its destabilization.

4.4 Conclusions

We have presented results from a modeling study of CFCs, ocean circulation, and

basal melting in the Ross Sea, Antarctica. Field data from a previous study [Smethie and

Jacobs, 2005] comprised of three different cruises were used to validate the modeled

temperatures, salinities, and CFC concentrations. Our model results suggest that plumes

of warm low CFC water that flow across the mouth of the RIS cavity may represent deep

81

Page 95: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

water (which has been isolated from the atmosphere and thus low in CFCs) that has

flowed onto the continental shelf and into the RIS cavity. This suggests the potential that

a markedly different mode exists in which warm MCDW directly enters the RIS cavity in

addition to the observed mode in which low-CFC water masses along the mouth of the

cavity originate from within the cavity.

82

Page 96: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 5

Ross Sea Primary Production and Sea Ice under Future Climatic Change Scenarios

Abstract

Average global temperatures are expected to rise 2-4 degrees Celsius in the next 50

years and winds are expected to increase in intensity and exhibit patterns different from the

present day. As these changes take place there will be unprecedented and unpredictable

impacts on ecological systems, particularly at high latitudes. Here we present results from a

modeling study of the effects of a range of expected changes in atmospheric temperature and

wind forcing on sea ice distributions and phytoplankton dynamics in the Ross Sea over the

next 50 years. Our results show that both increased temperatures and winds result in a year-

round reduction in sea ice in the Ross Sea. The winter sea ice maximum for the intermediate

temperature run (2.5°C increase in 50 years), intermediate wind run (1% increase per year for

50 years), and intermediate temperature+wind run will decline by 0%, 13%, and 11%,

respectively, by the end of the 50-year model run. The sea ice minimum will drop by 22%,

23%, and 26%, respectively, for these three runs. Associated with these reductions in sea ice

cover, annual primary production is predicted to decrease by 8%, increase by 16% and 8%

under altered atmospheric forcing in the 'temperature', 'wind', and 'temperature+wind' runs

respectively. These reflect substantial changes in primary production that may occur in the

next 50 years. Whereas the 'wind' run resulted in an increase in productivity, the most

notable predicted changes in primary production were in the low wind scenario (10% increase

in 50 years) in which peak production went up by 43% and the high wind scenario (a

doubling of winds in 50 years) in which peak production was delayed 20 days.

83

Page 97: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

5.1 Introduction

It is now well-established that global air temperatures are rising as a result of

increasing atmospheric greenhouse gas content and are likely to continue to increase for

the foreseeable future [Brohan et al, 2006]. Associated with this increase in temperature,

the number of severe weather systems as well as peak wind speeds are expected to rise

[Frauenfeld and Davis, 2003; Lucaritri and Russell, 2002; Simmonds and Keay, 2000]. Polar

regions have been identified as being uniquely vulnerable to change, and numerical

simulations of future climates by General Circulation Models (GCMs) agree in one

important respect: climate warming is predicted to begin earlier and manifest itself more

severely at the poles [Flato etal., 2000; Shindellet al, 1999; Stouffer et al, 1989]. For

example, high latitude precipitation is very likely to increase in part due to positive

feedbacks involving snow, sea ice, and other processes [Groisman et al., 2005]. The Arctic

has already undergone significant change, with surface air temperatures increasing faster

than almost anywhere on Earth, resulting in dramatic reductions in sea ice extent [Shimada

et al., 2006]. Although similarly striking changes in the Southern Hemisphere have mostly

been restricted to the Antarctic Peninsula region, it is expected that the rest of Antarctica

will soon exhibit significant changes due to global warming [IPCC, 2007].

The Southern Ocean currently plays an important role in the global carbon cycle,

accounting for a significant amount of the 2 Pg yr"1 of anthropogenic C 0 2 sequestered

from the atmosphere to the deep ocean [Caldeira andDujjy, 2000]. The uptake of C0 2 by

the ocean regulates the amount of this greenhouse gas in the atmosphere, thus helping to

control temperature and climate. At the same time, the rate of C 0 2 uptake by the ocean

is affected by changes in biogeochemical and physical oceanographic processes resulting

84

Page 98: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

from temperature increases. However there is limited understanding of the ocean's

physical and biological feedbacks to changing atmospheric C 0 2 [Sarmiento et al., 1998].

A number of feedbacks are expected to be expressed in the Southern Ocean.

Increasing temperatures and precipitation associated with warming should decrease the

density of surface waters, thereby increasing stratification and reducing the uptake of

atmospheric C 0 2 [Caldeira and Duffy, 2000]. Increased stratification will also reduce the

flux of nutrients from the deep ocean into the surface mixed layer, reducing

phytoplankton growth and sinking, hence the strength of the biological pump.

Additionally, rising ocean temperatures will decrease C0 2 solubility, reducing the ability

of the ocean to remove C 0 2 from the atmosphere. Similarly, continued intensification

and alteration of winds throughout the 21st century are likely to impact the ability of the

ocean to act as a sink for atmospheric C0 2 [Shindell and Schmidt, 2004]. These predictions

are consistent with observations showing that over the last 20 years, winds over the

Southern Ocean have increased by approximately 20% due to a positive trend in the

Southern Annual Mode (SAM) causing a poleward intensification of westerly winds south

of 45°S [Marshall, 2003; Marshall, 2007; Russell et al., 2006]. It is estimated that the

Southern Ocean C 0 2 sink has weakened by 0.08 petagrams C yr"1 between 1981 and 2004

because of the observed increase in winds and enhanced ventilation of C02-rich

subsurface waters \LeQuere et al., 2007]. This outgassing of C 0 2 dominates the variability

in the C 0 2 flux [Le Quere et al, 2007]. Conversely, increased upwelling during the positive

phase of the SAM has been shown to increase phytoplankton abundance by bringing

additional nutrients into surface waters [Lopenduski and Gruber, 2005] acting as a negative

feedback on C 0 2 outgassing.

85

Page 99: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Understanding the effects of climate change on distributions and annual cycles of

sea ice and primary productivity is critical. Antarctic polynyas (regions of open water

where ice is expected to be present) are sites of high primary productivity and serve as

key habitats for higher trophic level animals such as seals, penguins, sea birds and whales

[Lauriano et al, 2007]. One of the best ways to understand how climate change may alter

the sensitive physical and biological oceanographic systems of Antarctica is to simulate

these interactions using a mechanistic regional model. Here we present results of one

such study investigating the impacts of anticipated changes in air temperature and wind

speeds on sea ice concentrations and phytoplankton productivity in the Ross Sea sector

of the Southern Ocean.

5.2 Methods

5.2.1 The Physical Model

The physical model used in this study is the Polar Ocean Land Atmosphere and

Ice Regional (POLAIR) modeling system [Holland and Jenkins, 2001; Holland et al., 2003].

The ocean circulation component of the model comes from the Miami Isopycnic

Coordinate Ocean Model (MICOM) [Bleck and Smith, 1990]. MICOM has an embedded

mixed layer model based on a turbulent kinetic energy budget. The details of the

POLAIR model are summarized by Holland et al. [2003] and Reddy et al. [2007]. Briefly,

the Ross Ice Shelf is modeled using a viscous-sublayer parameterization describing

thermodynamic interactions at the interface between the ice shelf base and the ocean

surface [Holland and Jenkins, 1999]. The model has been modified to be compatible with

the surface topography of the base of the ice shelf [Holland and Jenkins, 2001]. The sea ice

model includes both dynamical and thermodynamical processes. Ice-ice interactions are

86

Page 100: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

treated using cavitating fluid rheology, whereby the ice pack does not resist divergence or

shear but does resist convergence. Where the sidewalls are boundaries with open ocean

(for the northern, eastern and western walls), a sponge zone of three grid cells is used.

For baroclinic fields, the restoring time scale for the exterior sponge cells is 10 days and

30 days for the interior cells. For barotropic fields, timescales of 1 day and 5 days are

used respectively.

The model grid uses a Mercator projection with cells being locally square and

becoming smaller moving southward in the domain. The average grid cell in the middle

of the Ross Sea Continental shelf is approximately 70 x 70 km. The model domain is

comprised of 40 grid cells in the longitudinal direction from 140°E to 120°W and 33 grid

cells in the latitudinal direction from 60°S to 85°S.

For the sea floor, a gridded version of the RIGGS data [Greischar et al., 1981] were

used south of 78°S, and ETOP05 was used to the north. The dataset used for the sub-

ice shelf cavity was derived from the Ross Ice Shelf Geophysical and Glaciological Survey

[Bentley andje^ek, 1981]. It was supplemented by ice front positions from the Antarctic

Digital Database [SCAR, 1993] and height measurements from the USCGC Polar Sea in

1994 [Keys et al, 1998]. The density of the modeled ocean was initialized using the

Hydrographic Atlas of the Southern Ocean Olbers et al. 1992]. A detailed explanation of

the datasets used is described by Holland et al. [2003]. For wind forcing, the model

computes geostrophic winds from the pressure gradient of daily sea level pressure data

fromNCEP [Kalnajetal, 1996].

87

Page 101: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

5.2.2 The Biological Model

The model for phytoplankton growth in the Ross Sea is derived from the

Coupled Ice, Atmosphere, and Ocean (CIAO) model of the same region \Arrigo et al,

2003]. Phytoplankton growth is modeled as a function of temperature, light, and nutrient

availability. Phytoplankton growth rate (ju [hr1]) is calculated using the temperature

dependent maximum growth rate (jumax [hr1]), and resource limitation (Resource^ Q), as

follows:

The temperature-dependent maximum growth rate is calculated as follows [Epp/ey, 1972]:

. , „0.063ir Mnax = Moe

where p0 is the net specific growth rate at 0°C (0.59 day"1) and Tis the temperature [°C].

Resource limitation is the minimum of light limitation (Enm) and iron limitation (Feu^)\

ResourceLim =mm(ELim,FeUm)

FeUm is calculated as

Fe Fe

K/iFe + Fe

where K1/2_FE [nM] is the temperature independent half-saturation constant for growth

(0.01 nM) and Fe [nM] is the concentration of iron [Arrigo et al, 1998; Monod, 1942]. Light

limitation is calculated from instantaneous photosynthetically available radiation, PAR

[jjiEin m 2 s"1], the photoacclimation parameter, EK [fjiEin m~2 s"1], and the photoinhibition

term, Inhib:

PAR

ELim=(\-e*« )• Inhib

Page 102: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

The photoacclimation parameter is a function of PAR and E & « , the maximum

photoacclimation parameter [McClain et al, 1996; Arrigo et al, 2003]

F

where E & a x is 30.0 [Arrigo et al, 2003] and

5 = exp[1.089-2.121og(i?imax)]

The photoinhibition term (Inhib), is calculated from PAR and the photoinhibition

threshold (100 uEin m"2 s"1) [Arrigo et al, 2003]:

Inhib = 1 -

20PAS

l-e m

l+5-108e m

Inhib remains near 1 between irradiances of 0 and the photoinhibition threshold.

Thereafter, Inhib drops quickly to 0.5 at the threshold and then down to 0 where it

remains as irradiance continues to increase.

PAR h\iEin m~2 s"1] for the mixed layer is calculated using Beer's law [McClain et al.,

1996]:

PARboltom=PARlope-K^

where PAR^^ [[jiEin m"2 s"1] is the PAR at the bottom of the layer, JPAR^ [[xEin m 2 s"1] is

the PAR at the top of the layer, and A% [m] is the layer thickness. The diffuse attenuation

coefficient for a given layer, KD [m4], is calculated as [McClain et al, 1996; Morel, 1988]:

KD = 0.027 + 0.0518(C/*/)0428

where Chi is the concentration of chlorophyll [mg chl m\ It then follows that the

average light for a given layer is:

PARto-PARhMom PARmidde - -

KDAz

89

Page 103: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Primary production (Produc. mg C m'3 hr'1) is calculated from the phytoplankton

growth rate (JA [hr1]) and concentration (Phyt, mg C m'3):

Produc = n • Phyt

The net source minus sink for tracers is a follows:

dPhyt _,. / , s — = n-Phyt -mortph (Phyt - Phyt^) - sinking

at

dFe _ /J • Phyt reminDet • Det

dt ~ C:Fe C:FeDet

dDet = (Phyt - i%?min ) • mortphyt - reminDst • Det - sinking

dt

where Det is the concentration of detritus, mortp^, is the mortality rate for phytoplankton

(0.01 hr'1), reminVet is the detrital remineralization rate (0.03 d"1) \Arrigo et al., 2003] and

QFe [mg C m'3:nM_Fe] is 5404.

5.2.3 Climate Change Scenarios

To better understand the potential impact of expected changes in environmental

conditions on the Ross Sea polynya ecosystem, we ran a number of model scenarios

including a control run and climate change runs whereby we slowly increased either the

air temperature (3 runs), wind speed (3 runs), or both (1 run) over the next 50 years. The

control run used NCEP reanalysis products for air temperatures and sea-level pressure

(which was used to compute geostrophic winds). To reduce variability in the model, the

same year of forcing data (1 January through 31 December 2000) was cycled for each of

the 50 years of the model runs. This was chosen as a representative year, in lieu of using

a climatology which mutes high wind events. We chose not to force our model with

90

Page 104: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

output from a global forecast model as we wanted to use simplified and targeted

atmospheric changes to assess the models sensitivity to such variations.

The temperature change runs were based on the global temperature rise projected

by the Special Report on Emissions Scenarios (SRES) emission scenarios. The low

temperature run was based on the mean of a range of SRES scenarios that assume

continued intensive reliance on fossil-fuels and emissions increases (Bl, AIT, B2, A1B

with 2100 C 0 2 concentrations of 600, 700, 800, and 850 ppm, respectively) [IPCC, 2007].

It produced a temperature increase of approximately 1.25°C over 50 years. The

intermediate temperature run was based on the A1FI SRES emission scenarios, the most

pessimistic IPCC scenario, with atmospheric C 0 2 concentrations in 2100 reaching 1550

ppm [IPCC, 2007]. This run increased temperature by 2.5°C in 50 years. The high

temperature run represented the upper extreme for temperature given by the Al FI SRES

emission scenario, increasing temperature by 5°C in 50 years. The atmospheric

temperatures for these three scenarios were increased through time as follows (Figure

5.1a):

AT (year) = 0.025 * (year-2005)

ATfyear) = 0.050 * (year-2005)

ATfyear) = 0.100 * (year-2005)

where ZlTis the increase in temperature above the value for the control run for a given

year of the model run.

Scenarios run to simulate the predicted intensification of peak wind speeds ranged

from a 10 to 100 % increase by 2050. It is not known how much polar wind intensities

may vary in the future. In order to determine an appropriate range for increased wind

91

Page 105: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

(a)

Temperature Increase Scenarios

— —

-

-5.0°C

2.5°C

-1.25X

(b)

2

2010 2020 2030 2040 2050

year

Wind Increase Scenarios

2000 2010 2020 2030 2040 2050

year

Figure 5.1 Increases in (a) temperatures and (b) winds over the 50 years of the model runs

intensities, we investigated wind speeds for a number of stations in the Ross Sea for a

single year. Between stations, the wind differed by between 10% and 300% in the austral

winter and by between 2% and 50% in the austral summer. Additionally, based on

estimates for the tropics, 1°C to 5°C of warming of the surface ocean would lead to

approximately a 5% to 25% increase in peak wind speeds [Emanuel, 1987; Emanuel, 2005].

Russell et al. [2006] used the SRES A1B scenario (a doubling of C 0 2 by 2100, stabilizing at

717) to force the Climate Model version 2.0 (CM2.0) and 2.1 (CM2.1), and predicted that

Southern Hemisphere westerlies will increase by 8% and 9%, respectively. This was used

92

Page 106: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

as the basis for the low wind increase run (with wind increased by 10%). (Climate Model

2.0 and 2.1 are the Princeton Geophysical Fluid Dynamics Laboratory's two coupled

climate models) The other two wind runs are high end sensitivity studies to see how the

system could respond to a 50% increase and a 100% increase in wind speeds. Wind

adjustment factors were calculated as a function of year as follows (Figure 5.1b):

wind'factor(year) — 0.002 * (year - 2000) + 1

windfactor(year) = 0.010 * {year- 2000) + 1

windfactor(year) = 0.020 * (year - 2000) + 1

For the different wind scenarios, the control run winds were multiplied by the wind

factors at each time step. A joint scenario was run by increasing both wind speed and

temperature as follows:

AT (year) - 0.05 * (year-2005)

windfactor(year)= 0.010 * (year- 2000) + 1

This temperature+wind scenario aims to quantify the combined effects of including

increases in both temperature and wind on sea ice distributions and phytoplankton

dynamics.

5.3 Results

5.3.1 Sea Ice

Over the 50 years of the model control run, mean sea ice concentrations over the

model domain oscillated between the summer sea ice minimum of 34-38% and the winter

sea ice maximum of 74-76%, with little interannual variability (Figure 5.2). For the

temperature run, (unless otherwise noted, the 'temperature run' will refer to the

intermediate temperature run and 'wind run' will refer to the intermediate wind run.) the

93

Page 107: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

85-

80-75-To­es 60

55-50-

45-40-35-

30-25-20-15-

(b) Intermediate Temperature

1 | j 1

| | | | j j | | l i l l l l l i l i i h l l i

111 li ill

1 I

1'

' 11 ' .1

1 1 !

1 In' n li M 'III l i ' 1 i'l

ill! 1 ml

'' P i li

1 1 1 1 i 1 1 ' i ill i III

| | 1 I I 1 ! 1 1 1 1 1 1 I 1 1 1 1

M l | mi it

ill 'I' III 1 I'l' ll 11 lllll

I''' 'ill"1! Iff 1 ,[J ' \ Is

lii'.,. iiiiiir!iii,i'1 limn M r i i> 1 !•' inn11'l'iuiiiiinri 1 i - ! | i i M j

2000 2005 2010 2015 2020 2025 2030 2035 2040 2045 2050

2000 2005 2010 2015 2020 2025 2030 2035 2040 2045 2050

year

2000 2005 2010 2016 2020 2025 2030 2035 2040 2045 2050

year

Figure 5.2 Sea ice concentration (averaged over the model domain) for the 50 years of the model runs for (a) control run (b) intermediate temperature run (c) intermediate wind run and (d) temperature+wind run

wintertime maximum varied between 73% and 75%, similar to the control run. However,

the summer minimum decreased from 36% at the beginning of the run to 28% after 50

years, a drop of 22%. Unlike the control run and the temperature run, the wintertime

maximum ice concentration for the wind run declined over the 50 year simulation,

dropping from 75% to 65%, going down by 13%. The summer minimum for this run

also fell, decreasing from 35% (2000-2005) to 27% (2045-2050), a decline of 23% during

the 50 years. Most of the decline occurred in the first 25 years, with the ice

concentrations for this run stabilizing after 2025. The temperature+wind run was

virtually identical to the wind run. The only difference was that the 50 year decrease in

the summer maximum was slightly less than that in the wind run. The similarity between

the wind run and the temperature+wind run means that the impact of temperature and

94

Page 108: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

78*S

140°E 160°E 180" 160"W 140"W 120°W 140*E IBO'E 180° 160"W 140"W 120*W 140"E 160°E 180* 160*W 140"W120*W 140*£ 160*E 180° 160*W 140eW120eW Longitude Longitude Longitude

60 80 100

Figure 5.3.1 Sea Ice Concentrations (light grey is open water; dark grey is the continent; light blue is the RIS) (a) control run (b) intermediate temperature ran (c) intermediate wind run (d) temperature+wind run

wind counteract each other. While the temperature run had less sea ice due to increased

stratification, the wind run had less ice due to additional ice advecting out of the area with

increased wind stress, despite the fact that higher winds mix the water column more

deeply and cool the mixed layer. So in the temperature+wind run, the winds bringing up

95

Page 109: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

I ̂ ggg ̂ ljllglj

-Q 66°S

140°E 160'E 180' 160'W 1400W120oW 140*E 160'E 180* 160*W 140'W120*W 140°E 160*E 180* 160°W 140eW120°W 140*E 160°E 180' 160*W 140'W120*W Longitude Longitude Longitude

0 20 40 60 80 100

Figure 5.3.2 Sea Ice Concentrations (light grey is open water; dark grey is the continent; light blue is the RIS) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run

colder water works to counteract the heating and stratification associated with increased

temperature.

It can be seen (Figure 5.3a and 5.3b) that there is slightly less sea ice in the Ross

Sea for model scenarios with increased temperature. The polynya starts to form in late

November and early December along the western side of the Ross Ice Shelf (RIS). The

96

Page 110: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

changes in atmospheric forcings do not dramatically influence the modeled formation of

the Ross Sea summertime polynya in December and January, causing it to open only a

little more quickly than it does in the control run. In December, the size of the polynya is

slighdy greater in the temperature run. The ice extent in the east is reduced as well, but

not as much as in the west. This small December polynya expands into a larger polynya

in the next few months. Breakthrough (whereby the polynya becomes contiguous with

the open ocean to the north) begins in January adjacent to the continent along the

western side of the model domain, approximately one month earlier than in the control

run. Additionally, the change in atmospheric forcing only slighdy retarded the closure of

the polynya in the autumn (April and May). Higher temperatures in the temperature run

do not substantially change sea ice cover from June through November.

The wind scenario (Figure 5.3c) shows a similar pattern, but amplified changes

compared to model results from the temperature run. The ice in the western region of

the model domain is dramatically reduced, with most of the ice receding from the

continental coast by 1 December, l ike the temperature run, breakthrough occurs by

January, whereas in the control run it does not occur until February. The ice cover in the

east is significandy reduced as early as December, a pattern that continues through

January and February. As the polynya closes in March and April, there is a narrower

band of ice along the continent in the east. In May, the polynya is fully closed in the

control run, but the northern extent of the ice is not as great in the wind run. Like in the

temperature run, there is an absence of ice along the coast of the continent in the west

and the ice does not reach as far to the north in the east. There are also changes noted in

the winter months of June through October, when increased wind speeds contribute to a

97

Page 111: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

wintertime polynya that is more persistently open, particularly in August. In August

through January, there is a reduction in sea ice concentrations compared to the control

run in the western section of the model domain (Figure 5.3c).

Changes due to increased temperature and winds are not additive, as can be seen

in the temperature+wind run results (Figure 5.3d). In December, the temperature+wind

ice extent is less than the control run and less than the temperature run, however it is

greater than the wind run. This means that temperature counteracts the effects of

increased winds alone in regards to sea ice (as described above). In January, there is only

slightly more ice in the temperature+wind run than in the wind run, and there continues

to be less ice than in the temperature run. By February and March the temperature+wind

run has slightly less ice than the wind run. The least amount of ice in the

temperature+wind run occurs in April and May as the polynya closes. In May, both the

temperature run and the temperature+wind run contain more open water than the other

runs in the western side of the domain next to the continent. In June and July, the

northern extent of the ice edge is lowest in the temperature+wind run, slightly less than

the wind run. In August, a wintertime polynya develops next to the RIS in the west,

similar to the temperature run. However, by September and October, the ice edge is

further north in the temperature+wind run than in the wind run, and continues as such

into November. In November, there are a couple of lower concentration patches of ice

in the center of the ice expanse in the temperature+wind run, similar to the wind run.

5.3.2 Ocean Temperatures

Ocean mixed layer temperatures in the temperature run reflect warming

atmospheric temperatures by the end of the 50 year run (Figure 5.4). In December, the

98

Page 112: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

(a) (b) (c) (d)

140-E 160°E 180* 160^140^120*™ 140°E 160°E 180° 160°W 140*W120"W 140*E 160'E 180* 160"W 140*W 120*W 140"E 160°E 180° 160*W 140°W 120BW

Longitude Longitude Longitude Longitude

- 2 - 1 0 1 2

Figure 5.4 Ocean mixed layer temperatures (dark grey is the continent; light blue is the RIS) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run

mixed layer temperatures in the control run (Figure 5.4a) in the polynya region are

approximately 0.5° - 1.0° cooler than in the temperature run (Figure 5.4b), with virtually

no differences elsewhere. This small temperature decrease is due to the region being

covered with sea ice during the preceding winter and early spring and thus was not

exposed to the warmer air temperatures. During this time, excess atmospheric heat

contributes to ice melt as opposed to ocean warming (as evidenced by the lower ice

concentrations in the region in December, Figure 5.3.1). This changes by March, when

surface waters have been ice-free for approximately three months and temperatures

throughout the polynya region are 1.0° - 1.5° warmer than in the control run.

99

Page 113: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Additionally, in the northern region of the domain, where the ocean remains ice-free all

year, the warm ocean waters extend further south than in the control run from December

through June.

Likewise, ocean mixed layer temperatures respond to the increasing winds in the

wind run by the end of the 50 years (Figure 5.4c). For this run in December, the mixed

layer in the polynya is slightly cooler than both the control run (approximately 0.5°C

cooler) and the temperature run (approximately 1°C cooler). Higher winds mix surface

waters more deeply, bringing up cooler waters from below in the polynya region and

spreading the surface heat over a larger volume of water. Surface temperature is

decreased by 0.75°C in the polynya region in March in the wind run. Again this is likely

due to the increased winds breaking down water column stratification in the polynya

region. Also notable is that warm surface waters (-1°C to +1°C) extend further south in

the east for the wind run. The explanation for this is that increased wind intensities,

particularly during the winter north of the ice edge, cause deeper mixing, in this case

bringing up heat from below (in the form of warm Circumpolar Deep Water, CDW) and

warming the mixed layer. In June, the polynya region is very similar in temperature to the

control run. Again in September there is a noticeable difference in the mixed layer

temperature in the northeast; once more likely due to the mixing up of warm CDW. The

polynya region is ice covered at this time of year and exhibits no difference from the

control run in this area.

In general, the ocean mixed layer is warmer in the temperature+wind run than in

the control run, but not as warm as the temperature run (Figure 5.4). In the east in

September and December, just north of the ice edge, the temperature of the

100

Page 114: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

temperature+wind run is cooler than in the wind run, and is more similar to the control

run. While the increased winds cool the polynya and counteract the warming effect of

increased atmospheric temperatures, in the northeast, increased winds make the ocean

warmer. In order of temperature in the polynya (December through March), the wind

run is the coolest, followed by the control run, the temperature+wind run, and then the

temperature run.

5.3.3 Primary Productivity

In all of the model simulations, the phytoplankton bloom began in mid-

November and primary production peaked very quickly in early December (Figure 5.5).

In the control tun, peak daily production (spatially averaged for all shelf waters less than

800 m depth, Figure 5.6) reached 2.66 g C m-2 d-1 during the week of 7 December

(Table 1). For the temperature run, the bloom developed more quickly but reached a

peak rate of primary production of only 2.18 g C m-2 d-1, significantly less than in the

control run. This is due to the smaller spatial extent of the bloom at its peak for the

temperature run (Figure 5.5). The wind run exhibited production rates that were lower

than in the control run, but the bloom covered a larger area near its peak (the week of 7

Table 5.1 Comparison of peak daily, mean daily, and annual primary production

Region All shelf waters (<800 m) [Arrigo et al. 2003] All shelf waters (<800 m), this study: control All shelf waters (<800 m), this study: temperature** All shelf waters (<800 m), this study: wind** All shelf waters (<800 m), this study: both**

Peak Daily Mean Daily Production Production* g C m"2 d"1 g C m"2 d"1

1.28 2.66 2.18 2.76 1.99

0.54 0.85 0.78 0.98 0.90

Annual Primary

Production g C m~2 yr"1

145 127 117 147 135

*Mean daily production is calculated for the time period 1 November through 1 April. **'temperature' and 'wind' refer to the intermediate increased run, respectively, 'both' refers to the increased temperature and wind run

Page 115: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

170°E 180' 170°W 160*W

Longitude

170°E 180e 170°W 160*W

Longitude

170BE 180" W W 160°W

Longitude

170eE 180' 170*W 160"W

Longitude

0.1 [gCm^d1]

Figure 5.5 Primary Production (dark grey is the continent) (a) control run (b) intermediate temperature run (c) intermediate wind run (d) temperature+wind run

December). As a result, the wind run had a peak daily production of 2.76 g C m-2 d-1,

the highest of the four model runs discussed here. The phytoplankton bloom in the

temperature+wind run developed more rapidly than in the other runs, yet had a peak

daily production of just 1.99 g C m-2 d-1, die lowest of the four model runs. Production

Page 116: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Date

Figure 5.6 Primary production time series

increased more quickly in this model run, similarly to the temperature run, because of a

slightly warmer mixed layer and increased stratification earlier in the season leading to

higher growth rates and biomass. The temperature+wind run had even higher

production early in the season than the temperature run because the increased winds in

the early spring mix nutrients into the surface waters, while at the same time, warmer

temperatures stratify the mixed layer.

Annual primary production is highest (for these 4 model runs) for the wind run at

147 g C m"2 yrA (for all shelf waters <800 m). Maximum daily rates of production are not

103

Page 117: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

as high as in some of the other runs, but this is compensated for by its larger spatial

coverage. The explanation for this is that increased winds mix the water column more

deeply in early spring prior to the bloom, bringing up nutrients (iron) from depth,

allowing the bloom to persist longer and expand over a larger area. Annual production is

similar for the control run and the temperature+wind run at 135 g C m 2 yr"1. This is in

part because increased temperatures and increased winds have a counteracting effect on

annual production because high winds tend to bring more nutrients to the surface while

higher temperatures increase stratification and reduce nutrient supply. This latter effect

can be seen clearly in the temperature run where annual production (117 g C m 2 yr"1) is

8% lower than in the control run.

5.3.4 Sensitivity to the Magnitude of Changes

Not surprisingly, sea ice extent, ocean mixed layer temperature, and primary

productivity were sensitive to changes in the magnitude of the air temperature and wind

speed changes during the 50 year simulation (Figure 5.7). Sea ice concentration (1

January) declined in response to increased temperature changes after 50 years. Loss of ice

was particularly evident in the western portion of the model domain (140°E-180°),

northwest of the main polynya (where ice concentrations dropped from approximately

40% to 0 between the control run and the low temperature run). The response to higher

temperatures was not large and it was fairly linear (loss of approximately 5% of ice

between runs as temperatures increased).

The change in January ice due to altered wind speed, however, was not linear.

The sea ice concentration in the low wind scenario was very similar to the control run.

However, in the intermediate wind scenario, sea ice concentrations decline substantially

104

Page 118: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

140"E 160"E 180* 160°W 140tWt20"W140''E 160"E 180' 160'W 140"W 120°W 140*E 160°E 180* 160SW 140°W 120°W

Longitude Longitude Longitude

0 20 40 60 80 100

78*S

140°E 160*E

140*E 160"E 180° leO'W 140*WI20*W14O'E 160*E 180* 160*W 140°W 120'W 140eE 160'E 180" 160°W 140*W 120*W

Longitude Longitude Longitude

17CE 180' 170°W 160°W

Longitude

170°E 180' 170"W 160°W 170«£ 180" 170*W 160*W

Longitude Longitude

0.1 [gcm-^d-1]

Figure 5.7 Model results from the low, intermediate, and high model scenarios for temperature and wind of (a) January sea ice concentrations (b) February ocean mixed layer temperatures and (c) December (peak) primary production (light grey is open water; dark grey is the continent; light blue is the RIS)

Page 119: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

compared to the control run. As the winds increased further in the high wind run, there

was very little change from the intermediate wind run. The size of the polynya under

different wind scenarios represents a balance between increased wind mixing bringing up

cooler waters (retarding ice melt) and increased wind advecting more ice out of the

region. Between the intermediate and high wind runs, this balance is such that the size of

the polynya changes very little (<5%).

Ocean mixed layer temperatures (1 February) warmed substantially in response to

increasing temperature by the end of the 50 year model runs. In the polynya region, there

was a significant increase (~ 0.75°C) between the low temperature run and the

intermediate temperature run, as well as between the intermediate and the high

temperature runs (~1°C). In die high temperature run, warmer waters (>2°C) in the

northern part of the model domain extended further south than in the other runs. In the

three different wind runs, changes in temperature were much smaller than in the

temperature runs. In the polynya region, the mixed layer temperatures were essentially

the same (~0.5°C) for die control run and die low and intermediate wind runs. For the

high wind run, sea surface temperatures were noticeably lower in die polynya, averaging

around -1°C. This decrease in temperature helps explain why there was not a substantial

reduction in sea ice from the intermediate to high wind runs.

Despite the changes in both sea ice concentration and mixed layer temperature at

higher atmospheric temperatures and wind speeds, peak primary productivity (on 7

December) changed very litde between the low, intermediate, and high air temperature

runs (2.64, 2.18, 2.38 g C m"2 d"1, respectively), but much more so in the wind runs (3.82,

2.76, and 2.48 g C m 2 d~\ respectively). Primary productivity in the low temperature run

106

Page 120: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

(2.64 g C m~2 d"1) was essentially the same as in the control ran (2.67 g C m"2 dA). The

intermediate temperature run had lower peak productivity (2.18 g C m"2 d"1) than the

control run, and the bloom was shifted slightly to the north. In the high temperature run,

the bloom was shifted even farther to the north than the intermediate run and peak

production was also slightly lower than in the control run (2.38 g C m 2 d"1).

There were much more pronounced changes in primary production for the

different wind runs. For the low wind run, there was significantly higher peak production

(3.82 g C m"2 d"1) than for any of the other runs. This is due to a balance of mixing

nutrients up to the surface early in the season and not breaking down stratification. In

the intermediate wind run, a reduction in maximum productivity was compensated for by

the larger extent of the bloom. In the high wind run, the bloom (when the other runs

peaked, 7 December) was not only more muted, but also covered a smaller area, leading

to substantially lower production at this time. The peak production was delayed until

later in the season (27 December), at which point the run reached a similar value of peak

production (2.48 g C m"2 d"1).

5.4 Discussion

5.4.1 Southern Ocean Sea Ice

During the IPCC [2007] study, a range of models was used to simulate Arctic and

Antarctic sea ice extent until the year 2100 for a number of different emission scenarios

[IPCC, 2007]. These models found that the reduction in ice extent in the Southern

Hemisphere was much greater in magnitude for July through September (approximately

2.5 x 106 km2 of 17.7 x 106 km2, or 14% in 50 yr) than in January through March

(approximately 1.0 x 106 km2 of 3.9 x 106 km2, or 25% in 50 yr), but lower on a

107

Page 121: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

percentage basis. Like the IPCC results, our model predicts similar winter changes in ice

extent for the wind and temperature+wind runs (13%, and 11%, respectively) and less for

the temperature run (0%). Additionally, our model predicts the same percent ice loss

during the summer (a drop of 22%, 23%, and 26% for temperature, wind and

temperature+wind runs) as the IPCC change in ice extent from January through March.

There are some differences between the models that make direct comparisons somewhat

difficult. First, the IPCC models are not of high enough resolution to capture the Ross

Sea polynya, so for the most part, all of the changes in sea ice extent are due to changes

of the northern ice extent. Thus, any ice loss in the polynya region would not be

captured by the IPCC models. Second, the IPCC models contain a higher percentage of

open ocean (north of the ice extent) exposed to warmer temperatures year round.

Warming of the air overlying the surface of the ice, which is already very cold, contributes

little to ice melt. A much larger component of ice melt is attributable to warming of

ocean water. Seawater has a much lower albedo than sea ice, and thus, warms more than

ice does from solar radiation. Since there is a bigger impact of ocean heat than air

temperatures on sea ice, a larger percent of model domain that is open ocean exposed to

heating could contribute to greater ice melt in the IPCC results. However this is highly

dependent on ocean circulation patterns.

In addition to affecting heating of ocean waters below, changes in ice cover

affects air-sea gas exchange. A reduction in sea ice extent, as occurs through the

formation of wintertime polynyas, has implications for the exchange of C 0 2 between the

ocean and the atmosphere [Sweeney et al, 2000]. A wintertime polynya forms immediately

north of the western side of the RIS during most of the runs. The temperature run

108

Page 122: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

wintertime polynya is largest in August, and is similar in size to the polynya in the

temperature+wind run (also in August). By September, the polynya in the temperature

run is still approximately the same size. In the wind run, the polynya is completely closed

in September and the polynya in the temperature+wind run shrinks noticeably. However,

by October, it is the wind run that has the most open water in that region. The high

winds that keep the polynya open in the winter also lead to high rates of gas exchange.

Compared with ice covered waters, gas fluxes per unit area in the open water of these

winter polynyas can be very large. Since the polynyas open up in August and their area is

relatively small, they do not contribute gready to the overall annual net flux of gases

(between the atmosphere and the ocean) in the Ross Sea. The annual net flux of C 0 2 is

as high as -2.9 mol C m"2 in the Ross Sea, however, most of this flux occurs between

December and March \Arrigo and Van Dijken, 2007]. In the autumn and winter, die

release of C 0 2 from the ocean into the atmosphere is 0.05 to 0.09 mol m"2 yr"1 \Arrigo and

Van Dijken, 2007]. Recent evidence suggests that Antarctic continental shelves have high

biological drawdown of C 0 2 \Arrigo et al, submitted] and that Southern Ocean storage of

anthropogenic C 0 2 is likely higher than was previously believed \LJJ Monaco et al, 2005].

However, there is litde evidence that wintertime polynyas play a significant role in

Southern Ocean C uptake.

5.4.2 Arctic and Antarctic Differences

In the IPCC model runs mentioned previously, ice loss is predicted to be greater

in the Northern Hemisphere than in the Southern Hemisphere. In some projections,

late-summer ice in the Arctic is predicted to disappear almost entirely by the latter part of

the 21st century. Some models predict a tipping point in the Arctic in as little as 40 years

109

Page 123: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

[Lindsay and Zhang, 2005; Winton, 2006] at which time sea ice will disappear completely

during the summer. This is expected to occur in large part due to positive feedbacks with

the albedo affect. Because the open water of the ocean has a much lower albedo than the

ice, the more waters become ice free, the more rapidly warming occurs. It is likely that

the Antarctic has a similar sort of tipping point, after which summer sea ice significantly

drops and then verges on disappearing in a region. Although our model shows does not

suggest that a tipping point will occur in the Ross Sea within the next 50 years, we cannot

rule out the possibility that such a tipping point may occur after that time. Unlike the

Arctic, where sea ice is centralized over the pole, sea ice in the Antarctic encircles the

continent and is found at very different latitudes that are exposed to varying localized

climates. These regional differences could lead to sea ice tipping points happening at

different times around the continent. It has already been observed that ice loss is much

greater along the Antarctic Peninsula, the northernmost part of Antarctica.

Another reason that Antarctic sea ice concentrations have changed little

compared to the Arctic, and are predicted to change to a smaller extent in the future, may

be due to the sluggish warming of the Southern Ocean compared to the Arctic [IPCC,

2007]. In the Ross Sea, sea ice concentrations associated with the polynya have been

shown to be affected by ocean temperatures as well as winds \Reddy et al, 2007]. Because

the ocean is such a large reservoir, there is a lag between changes in atmospheric

temperature and the same degree of warming in the ocean. However, the magnitude of

this lag is likely to differ for these two polar regions due to their large differences in

bathymetry. In particular, the Arctic is generally shallower (average depth 1000 m), with

much of the basin occupied by continental shelf. This leaves it more susceptible to

110

Page 124: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

heating by warmer air temperatures than the Southern Ocean, which is much deeper

(average depth 4000-5000 m). Mixing in the Southern ocean throughout the year spreads

the heat throughout a deep water column. Furthermore, the Antarctic continental

shelves remain ice covered throughout much of the year, limiting their exposure to

atmospheric influences. Model results for the increased wind runs show a signature of

cooling in surface waters in die polynya (however, to the northeast, it does show

warming; this is caused by the mixing up of warm CDW). If the scenario was run for the

Arctic, where waters on the continental shelves are ice free for a longer time and exposed

to solar radiation earlier in the season, the increased wind runs may not have exhibited

the same degree of cooling. Similarly, our increased temperature runs likely would have

exhibited greater warming over the continental shelf since they were ice free longer as

those in the Arctic.

In addition to warmer atmospheric temperatures, declines in Arctic ice cover have

been attributed in part to the advection of warm water into the Arctic ocean from the

Pacific [Maslowski et al., 2001; Shimada et al., 2006] and Atlantic [Dickson et al., 2000; Steele

and Boyd, 1998]. Unlike the Arctic Ocean, a shallow basin surrounded by land with warm

water inflows, the Antarctic is bordered by deep continental shelves and the Antarctic

Circumpolar Current (ACQ. We have not seen as much change in Antarctic coastal

waters as compared to Arctic waters in part because we have made fewer observations

there.

5.4.3 Response of Biological Systems to Changes in Physical Systems

Our peak daily production rates agree well with Arrigo et al. [2003], who estimate

that peak daily production is 1.28 g C m"2 d"1 for all shelf waters less than 800 m (Table 1).

I l l

Page 125: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Annual production oi Arrigo et al. [2003] is 145 g C m 2 yr"1, which is in the same range as

our estimates. The timing of the spring blooms also agree well (Figure 5.6), with daily

production peaking in early December in both the study by Arrigo et al. [2003] and ours.

One difference is that after the peak of the spring bloom, our model predicts that primary

production will drop off more quickly in December than was reported by Arrigo et al.

[2003]. In both studies, primary production declines steadily from January until late

March. The peak and then drop-off in our model estimates reflect the boom and bust

nature of the modeled bloom in the central Ross Sea. This depletes nutrients in the

bloom region and production rapidly declines. Nutrients are not depleted on the margins

of the original bloom, allowing for production to take place in these regions. This results

in spatially averaged daily production staying fairly steady throughout the summer, as the

bloom expands outward into nutrient-rich waters.

In general, warmer waters will increase phytoplankton growth rates, and it is

expected that levels of production will be higher. However, we do not see this reflected

in mean daily production or the annual production for all of the temperature runs, which

exhibits rates that are lower than those in the wind and temperature+wind runs. It is

notable that mean daily production is higher in late November for the high and

intermediate temperature run. The increase in production early in the season is

outweighed in part by slightly lower peak daily production and lower daily production

throughout the season thereafter. The changes from the control run to all three

temperature runs are not very large (<20%), perhaps due to the limited changes in ocean

temperatures (<0.5°C change) in the polynya region this early in the season. Modeled

ocean temperatures (for the intermediate temperature run) do not substantially increase

112

Page 126: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

above the control run (by more than 0.5°C change) until January and February, with

considerably warmer water (1.5°C change) in the polynya region by March (Figure 5.4).

In the intermediate wind run, ocean temperatures are slightly cooler in the

polynya in both the early part of the season (December, Figure 5.4) and later on (March,

Figure 5.4). The effect of this on the bloom is that it does not reach the high levels of

peak production predicted in the control run, although its spatially averaged peak daily

production is greater due to a larger bloom extent. This greater spatial extent of the

bloom is attributable to increased winds in the early spring mixing nutrients into the

surface waters creating a more muted bloom but over a larger area. At the time of peak

daily production and thereafter in the season, this intermediate increase wind run has the

highest daily production levels of any of the four run discussed here (control run,

intermediate temperature run, intermediate wind run, temperature+wind run). It follows

that annual production is the highest for the intermediate wind run than for any of the

four runs (147 g C m 2 yr"1). This is due to deeper mixing increasing nutrient supply to

the surface waters outweighing cooler mixed layer temperature in the polynya.

It has already been observed that in some marine systems, shifts in the ranges and

abundance of algal, plankton, and fish populations are associated with rising water

temperatures and related changes in ice cover, salinity, oxygen levels and circulation

[Beaugrandetal, 2002; Chave^et al, 2003; Richardson andSchoeman, 2004]. The Arctic has

experienced a 27.5 Tg C yr4 increase in primary productivity over the past five years due

to both a decrease in the summer minimum sea ice extent and the longer ice-free growing

season \Arrigo et al, submitted]. In our model results, we do not see a marked change in

primary production in the Ross Sea, in large part because there is not a significant

113

Page 127: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

reduction in the summer minimum ice extent or extension of the ice-free growing season.

Our modeled productivity develops very quickly in early December and then dies out

quite rapidly by the end of the month, and then stays at a moderate level until early April.

Due to the long tail at the end of the bloom, productivity would likely increase if the

polynya were to remain open longer, as has been observed in the Arctic \Arrigo et al,

submitted], unless the effect of increased wind causes a drop in production, as we predict

in our high wind run. If winds are not strong enough to inhibit the bloom, an extended

growing season would prolong the bloom if nutrients were not drawn down completely

and there was still an adequate amount of light. If nutrient were depleted, an extension of

the ice-free growing season may allow enough time for remineralization, fueling a

secondary bloom. However, we do not see a dramatic change in the timing of polynya

closure in any of our simulations. In order to prolong the phytoplankton bloom, the ice-

free growing season would need to be significantly longer than what we have predicted.

If future climatic changes alter nutrient supply to the summertime polynya (such as by

higher temperature increasing stratification or greater wind speeds deepening mixing), in

addition to lengthening the ice-free growing season, we could expect to see an increase in

production. On the other hand, the bloom may be shortened if the higher peak levels of

production associated with warmer temperature exhaust nutrients sooner.

5.5 Conclusions

Model results suggest that with increasing air temperature, sea ice extent will be

reduced slightly, mixed layer ocean temperatures will warm by a few degrees in the

polynya region, and peak primary production will decrease slightly. With increasing in

wind speeds, sea ice concentrations will drop slightly with a small (10%) increase in

114

Page 128: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

winds, and more substantially with a moderate (50%) or large (100%) increase. Ocean

temperature in the polynya will stay almost the same in response to a small increase in

winds and then drop slighdy with a larger increase. Increased winds lead to a more

muted bloom that covers a larger area except when the model is run with very high wind

speeds (100% increase) at which point the bloom is delayed almost a month.

It is obvious that the scenarios run in this study are simplifications of changes that

are likely to actually occur with climate change. It is expected that both air temperatures

and winds speed will increase, but also that the pattern of winds may be altered [Russell et

a/., 2006], which may play a role in determining sea ice concentrations and primary

productivity in the Ross Sea. Additionally, in the future, there will be changes not

discussed here such as physical conditions and pressures on Antarctic ecosystems. For

example, we did not include atmospheric changes such as clouds or ocean changes such

as advection of different water masses from outside of die domain. Ecosystem pressures

include ocean acidification, which is expected to have negative impacts on calcium

carbonate shell-forming organisms such as pteropods.

We predict that changes in sea ice and primary production will be smaller than

those that have been observed in the Arctic, with no indication of a tipping point during

the 50 year model run. This is likely the result of differences in topography between the

two regions, particularly the greater depth of the Antarctic shelves and basins, the

relatively enclosed boundaries and warm water inflow into the Arctic Ocean, and the

regional differences in climate that characterize the Antarctic continent.

115

Page 129: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

CHAPTER 6 Conclusions

The research presented here has helped to paint a picture of ocean circulation in

the Ross Sea using techniques involving quantities as varied as chlorophyll, sea ice, and

oceanic CFCs. From these studies we corroborate earlier findings that warm MCDW

moves onto the shelf and through deep mixing in the winter, inhibits ice growth and thus

contributes to the polynya opening in the summer. This water mass is also studied using

CFCs as a tracer, with model results showing that plumes of warm, low-CFC water enter

the Ross Ice Shelf (RIS) cavity year-round, originating from off of the shelf. This mode

in which MCDW directly enters the ice shelf cavity is observed some of the time in field

observations. Its importance should not be undervalued, as the presence of this warm

water could have a destabilizing effect on the RIS. Finally, patterns observed in an

analysis of satellite chlorophyll, as well as supporting modeling work, show that the

geostrophic nature of the flow of polar waters in the Ross Sea pushes water onto the

shelf and around shoals, with a connection from the bottom of the water column to the

top that leads to a signature of bathymetry on these intrusions in surface waters.

This dissertation represents a significant contribution to our understanding of

the oceanographic system in the Ross Sea. Specifically, it has expanded our scientific

knowledge of:

• MCDW and Phytoplankton Blooms. Circumpolar Deep Water (CDW)

is a water mass located offshore of the Antarctic continent that is warm,

high in nutrients, and low in CFCs. Contrary to what has been suggested

116

Page 130: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

previously, model results show that Modified Circumpolar Deep Water

(MCDW) is not entrained into the surface layer in the summer and

contribute nutrients to the phytoplankton bloom at this time of year.

Rather, MCDW appears to mix into the surface layer only during the winter,

minimizing its potential impact on phytoplankton dynamics.

• Extent of the Ross Sea Phytoplankton Bloom. Highly productive

waters of the Ross Sea are confined to the continental shelf by the

geostrophic flow of offshore waters that have low chlorophyll and iron

concentrations. Interannual variability in the maximum extent of the bloom

may result from a balance between the availability of iron from deep mixing

on the shelf and the strength of geostrophic circulation along the shelf

break. Previous studies suggested that biological activity is enhanced by the

movement of nutrient-rich CDW onto the continental shelf. Instead, we

suggest that there is a reorganization of surface flow rather than

entrainment of high iron containing intermediate waters into the surface,

and diat instead of offshore waters contributing to biological production (by

entrainment of deep water high in nutrients), they likely inhibit it (by

intrusion of surface water low in nutrients).

• Formation and Evolution of the Ross Sea Polynya. In agreement with

current scientific understanding, model results suggest that both thermal

and mechanical sea processes contribute to the formation of the Ross Sea

polynya. However, model results show that wind-forced sea ice motion

plays the primary role in polynya formation, with entrainment of warm

117

Page 131: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

MCDW into the surface layer during the winter (when it retards the growth

of ice) playing a relatively minor role.

• Residence Time & Basal Melting. By modeling CFC distributions in the

Ross Sea, we estimated the residence time of water in the RIS cavity to be

2.2 years. More importantiy, the model predicts a mode of circulation,

observed occasionally in field data, in which MCDW directly enters the RIS

cavity and contributes to a basal melting rate of 10 cm yr"1. This flow of

warm water into the RIS cavity could have grave impacts on future melting

of the ice shelf, especially in light of increases in water temperature expected

in response to global warming.

• Future Projections. Model runs projecting 50 years into the future under

scenarios of increased temperatures and winds reveals that there is slighdy

less sea ice both during the austral summer and year-round (a decrease in

the summer minima ice extent by 14% and 22%, respectively). The

wintertime polynya along the edge of the RIS is larger in both scenarios.

Primary production is forecast to change in magnitude and timing, but only

slighdy. Increasing temperatures leads to a stronger peak bloom that ends

more quickly due to nutrient limitation. On the other hand, increased winds

lead to a more muted bloom. It is also possible diat such changes in

physical conditions will lead to a shift in plankton community structure and

could have more widespread ecological impacts than solely altering the

magnitude of the bloom.

118

Page 132: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

In light of the recent changes observed around the globe and those that are

expected in the future, expanding our understanding of important and vulnerable Earth

systems is critical. The Ross Sea is an ecologically active region, with one of the largest

polar phytoplankton blooms, and with a complex web of upper trophic level organisms

that it supports. As a polar region, it is expected to be especially vulnerable to climate

warming. In our study of the impact of winds and temperature on sea ice and

productivity, we do not predict major changes in the next 50 years. However, a close

watch will have to be kept on the RIS and its melt rates and stability. Our current

understanding of the mass balance of the ice shelf is very limited. If temperatures warm

enough, it could have catastrophic consequences for the ice shelf. One avenue for its

destruction could be an input of warm water into the cavity, originating from offshore as

CDW. However, if this flow of warm water were to intensify in volume or frequency in

the future under warming conditions, it could contribute substantially to the melt of the

ice shelf. The ice shelf, because it is floating, will not contribute significantly to sea level

rise if it melts. However, if it disintegrates in part or entirely, it will no longer inhibit

continental ice from flowing into the ocean, and thus would alter sea level dramatically.

In the research presented here, we add to our collective understanding of the

Ross Sea system. Hopefully, through this work, we have made it easier for scientists in

the future to track ocean circulation, evaluate ice shelf melting, study polynya formation,

assess phytoplankton and ecosystems, and understand causes and predict consequences

of future environmental change, both in the Ross Sea, and elsewhere in the world.

119

Page 133: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

References

Ackley, S. F., J. L. Bengtson, P. Boveng, M. Castellini, K. L. Daly, S.Jacobs, G. L. Kooyman, J. Laake, L. Quetin, R. Ross, D. B. Siniff, B. S. Stewart, I. Stirling, J. Torres and P. K. Yochem (2003), A top-down, multidisciplinary study of the structure and function of the pack-ice ecosystem in the eastern Ross Sea, Antarctica, Polar Rec, 39, 219-230.

Ainley, D. G., E. D. Clarke, K. R. Arrigo, W. R. Fraser, A. Kate, K. J. Barton and P. R. Wilson (2005), Decadal-scale changes in the climate and biota of the Pacific sector of the Southern Ocean, 1950s to the 1990s, Antarct, Sci., 17(2), 171-182, doi:10.1017/S0954102005002567.

Ainley, D. G., R. C. Wood and W. J. L. Sladen (1978), Bird life at Cape Crozier Ross Island, Wilson Bull., 90, 492-510.

Anderson, J. B., C. F. Brake and N. C. Myers (1984), Sedimentation on the Ross Sea Continental Shelf, Antarctica, Marine Geol., 57, 295-333.

Aoyanagi, M. and K. Ueda (1989), Report of the First International Conference on Penguins Dunedin New Zealand, Antarc. Rec, 33, 88-95.

Arrigo, K. R, G. R. DiTullio, R. B. Dunbar, D. H. Robinson, M. L. Van Woert, D. L. Worthen and M. P. Lizotte (2000), Phytoplankton taxonomic variability in nutrient utilization and primary production in the Ross Sea,/. Geophys. Res., 105, 8827-8846.

Arrigo, K. R., D. H. Robinson, R. B. Dunbar, A. R. Leventer and M. P. Lizotte (2003), Physical control of chlorophyll a, POC, and TPN distributions in the pack ice of the Ross Sea, Antarctica,/. Geophys. Res., 108(C10), 1007-1021, doi:10.1029/2001JC00113.

Arrigo, K. R., D. H. Robinson, D. L. Worthen, R. B. Dunbar, G. R. DiTullio, M. L. Van Woert and M. P. Lizotte (1999), Phytoplankton community structure and the drawdown of nutrients and C 0 2 in the Southern Ocean, Science, 283, 365-367.

Arrigo, K. R. and G. L. Van Dijken (2007), Interannual variation in air-sea C 0 2 flux in the Ross Sea, Antarctica: a model analysis, Journal of GeophsicalResearch-Part C-Oceans, 112,16.

Arrigo, K. R. and G. L. van Dijken (2003a), Impact of iceberg C-19 on Ross Sea primary production, Geophys. Res. Lett., 30(16), 1836, doi:10.1029/2003GL017721.

Arrigo, K. R. and G. L. van Dijken (2003b), Phytoplankton dynamics within 37 Antarctic coastal polynya systems,/. Geophys. Res., 108(C8), 3271, doi:10.1029/2002JC001739.

120

Page 134: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Arrigo, K. R. and G. L. van Dijken (2004), Annual changes in sea-ice, chlorophyll a, and primary production in the Ross Sea, Antarctica, Deep Sea Res. II, 5/(1-3), 117-138, doi:10.1016/j.dsr2.2003.04.003.

Arrigo, K. R., G. L. van Dijken and S. Pabi (submitted), The impact of a shrinking Arctic ice cover on marine primary production, Geophys. Res. Lett.

Arrigo, K. R., A. M. Weiss and W. O. Smith (1998), Physical forcing of phytoplankton dynamics in the southwestern Ross Sea,/. Geophys. Res., 103,1007-1021.

Arrigo, K. R, D. L. Worthen and D. H. Robinson (2003), A coupled ocean-ecosystem model of the Ross Sea: 2. Iron regulation of phytoplankton taxonomic variability and primary production,/. Geophys. Res., 108(CT), 3231, doi:10.1029/2001JC000856.

Arrigo, K. R., D. L. Worthen, A. Schnell and M. P. Lizotte (1998), Primary production in Southern Ocean waters,/. Geophys. Res., 103(C8), 15587-600.

Assmann, K., H. H. Hellmer and A. Beckmann (2003), Seasonal variation in circulation and water mass distribution on the Ross Sea continental shelf, Antarct. Set., 15(1), 3-ll,doi:10.1017/S0954102003001007.

Bates, N. R., D. A. Hansell, C. A. Carlson and L. I. Gordon (1998), Distribution of C 0 2

species, estimates of net community production, and air-sea C 0 2 exchange in the Ross Seapolynya,/ Geophys. Res., 103, 2883-2896.

Beaugrand, G., P. C. Reid, F. Ibanez, J. A. Lindley and M. Edwards (2002), Reorganization of North Atlantic marine copepod biodiversity and climate, Science (Washington D C), 296(5513), 1692-1694.

Bentley, C. R. and K. C. Jezek (1981), RISS, RISP AND RIGGS - Post-IGY glaciological investigations of the Ross Ice Shelf in the US Programme, J.KSocN.Z., 11(A), 355-372.

Bindoff, N. L., G. D. Williams and I. Allison (2001), Sea-ice growth and water-mass modification in the Mertz Glacier polynya, East Antarctica, during winter, Ann. Glaciol, 33, 399-406.

Bleck, R. and L. T. Smith (1990), A wind-driven isopyenic coordinate model of the North and equatorial Atlantic Ocean. 1. Model development and supporting experiments, /. Geophys. Res., 95, 3273-3285.

Blumberg A. F. and G. L. Mellor (1987), A Description of a Three-dimensional Coastal Ocean Circulation Model, in Three Dimensional Coastal Ocean Models, edited by N. S. Heaps, pp. 1-16, AGU, Washington, D.C.

Brohan, P., J. J. Kennedy, I. Harris, S. F. B. Tett and P. D.Jones (2006), Uncertainty estimates in regional and global observed temperature changes: A new data set from 1850, Journal of'Geophysical'Research'Atmospheres, 111, D12106.

121

Page 135: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Bromwich, D. H., J. F. Carrasco and R. Y. Tzeng (1993), Hemispheric atmospheric variations and oceanographic impacts associated with katabatic surges across the Ross Ice Shelf, Antarctica,/. Geophys. Res., 98,13045-13062.

Bromwich, D. H. and D. D. Kurtz (1984), Katabatic wind forcing of the Terra Nova Bay polynya,/. Geophys, Res., 89, 3561-3572.

Bromwich D. H., Z. Liu, A. N. Rogers and M. L. Van Woert (1998), Winter atmospheric forcing of the Ross Sea Polynya, in Ocean, Ice and Atmosphere: Interactions at the Antarctic Continental Margin, Antarct. Res. Ser., vol. 75, edited by S. S. Jacobs and R. F. Weiss, pp. 101-133, AGU, Washington, D.C.

Bromwich, D. H., A. J. Monaghan, A. N. Rogers, M. L. Van Woert and K. R. Arrigo (2001), Atmospheric forcing of the Ross Sea polynya during summer, The 6th

Conference on Polar Meteorology and Oceanography and the 11th Conference on Interaction of the Sea and Atmosphere, Am. Meteorol. Soc, San Diego, Calif., 14-18 May.

Bu, X. and M.J. Warner (1995), Solubility of Chlorofluorocarbon-113 in Water and Seawater, Deep Sea Res. I, 42(7), 1151-1161.

Buffoni, G., A. Cappelletti and P. Picco (2002), An investigation of thermohaline circulation in Terra Nova Bay polynya, Antarct. Sci., 14(1), 83-92, doi:10.1017/S0954102002000615.

Caldeira, K. and P. B. Duffy (2000), The role of the Southern Ocean in uptake and storage of anthropogenic carbon dioxide, Science, 287(5453), 620-2.

Chavez, F. P., J. Ryan, S. E. Iiuch-Cota and N. C. Miguel (2003), Climate: From anchovies to sardines and back: Multidecadal change in the Pacific Ocean, Science (Washington D C), 299(5604), 217-221.

Clarke A. and N. M. Johnston (2003), Antarctic marine benthic diversity, in Oceanography and Marine Biology, vol. 41, edited by R. N. Gibson and R. J. A. Atkinson, pp. 47-114, Aberdeen University Press, London.

Coale, K. H., R. M. Gordon and X. J. Wang (2005), The distribution and behavior of dissolved and particulate iron and zinc in the Ross Sea and Antarctic circumpolar current along 170 DGW, Deep Sea Res., Parti, 52, 295-318.

Cochlan, W. P., D. A. Bronk and K. H. Coale (2002), Trace metals and nitrogenous nutrition of Antarctic phytoplankton: experimental observations in the Ross Sea, Deep Sea Res., Part II, 49, 3365-3390.

Comiso, J. C , C. R. McClain, C. W. Sullivan, J. P. Ryan and C. L. Leonard (1993), Costal Zone Color Scanner pigment concentrations in the Southern Ocean and relationships to geophysical surface features,/. Geophys. Res., 98, 2419-2451.

Page 136: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Court, G. S., L. S. Davies, S. Focardi, R. Bargargli, C. Fossi, C. Leonzio and L. Marili (1997), Chlorinated hydrocarbons in the tissues of South Polar Skuas (Catharacta maccormicki) and Adelie Penguins (Pygoscelis adeliae) from Ross Sea, Antarctica, Environ. Poll., 97, 295-301.

Darby, M. S., A. J. Willmott and T. A. Somerville (1995), On die influence of coastline orientation on the steady state width of a latent heat polynya,/. Geophys. Res., WO, 13625-13633.

Dare, R. A. and B. W. Atkinson (2000), Atmospheric response to spatial variations in concentration and size of polynyas in the Southern Ocean sea-ice zone, Bound-Layer Meteorol., 94(1), 65-88, doi:10.1023/A:1002442212593.

Davis, A. M.J. and R. T. McNider (1997), The development of Antarctic katabatic winds and implications for the coastal ocean,/. Atmos. Set., 54,1248-1261.

Deliberty, T. L. and C. A. Geiger (2005), Temporal and Regional Variations in Sea Ice Thickness in the Ross Sea during 1995 and 1998 , 8th Conference on Polar Meteorology and Oceanography, Am. Meteorol. Soc, San Diego, Calif., 9-13 January.

Dery, S. J. and L. B. Tremblay (2004), Modeling the Effects of Wind Redistribution on the Snow Mass Budget of Polar Sea Ice,/. Phys. Oceanogr., 34(1), 258-271, doi:10.1175/1520-0485(2004)034<0258:MTEOWR>2.0.CO;2.

Dewitt H. H. (1970), The Character of the mid water fish fauna of the Ross Sea Antarctca, in Antarctic ecology, vol. 1, edited by M. W. Holdgat, pp. 305-314, Academic Press, London.

Dickson, R. R., T.J. Osborn, J. W. Hurrell, J. Meincke, J. Blindheim, B. Adlandsvik, T. Vinje, G, Alekseev and W. Maslowski (2000), The Arctic Ocean response to the North Atlantic Oscillation, J.Clim., 13(15), 2671-96.

Dinniman, M. S., J. M. Klinck and W. O. Smith (2003), Cross-shelf exchange in a model of the Ross Sea circulation and biogeochemistry, Deep Sea Res., Part II, 50(22-26), 3103-3120, doi:10.1016/j.dsr2.2003.07.011.

Edwards, R. and P. Sedwick (2001), Iron in East Antarctic snow: Implications for atmospheric iron deposition and algal production in Antarctic waters, Geophys. Res. Lett., 28, 3907-3910.

Emanuel, K. (2005), Increasing destructiveness of tropical cyclones over the past 30 years, Nature, 436(7051), 686-8.

Emanuel, K. A. (1987), The dependence of hurricane intensity on climate, Nature, 326(6112), 483-5.

Page 137: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Eppley, R. W. (1972), Temperature and Phytoplankton Growth in the Sea, Fish. Bull., 70(A), 1063-1085.

Erickson, A. W., R. J. Hofman, R. J. Oehlensc, J. Otis and D. Kuehn (1971), Seal Population Studies in the Ross Sea, Antarct. J. U. S., 6, 98-99.

Fichefet, T. and H. Goosse (1999), A numerical investigation of the spring Ross Sea polynya, Geophys. Res. Lett., 26,1015-1018.

Flament, P., L. Armi and L. Washburn (1985), The evolving structure of an upwelling filament,/. Geophys. Res., 90, 11765-11778.

Flato, G. M., G. J. Boer, W. G. Lee, N. A. McFarlane, D. Ramsden, M. C. Reader and A. J. Weaver (2000), The Canadian Centre for Climate Modelling and Analysis Global Coupled Model and its Climate, Clim. Dyn., 16(6), 451-467.

Flato, G. M. and W. D. Hibler (1992), Modeling pack ice as a cavitating fluid,/. Phjs. Oceanogr., 22, 626-651.

Frauenfeld, O. W. and R. E. Davis (2003), Northern Hemisphere circumpolar vortex trends and climate change implications, Journal of Geophysical Research, 108, ACL7-1.

Gallee, H. (1997), Air-sea interactions over Terra Nova Bay during-winter: simulation with a coupled atmosphere-polynya model,/. Geophys. Res., 102,13835-13849.

Gordon, A. L. and J. C. Comiso (1988), Polynyas in the Southern Ocean, Sri. Am., 258, 10-11.

Gordon, H. R , D. K. Clark, J. W. Brown, O. B. Brown, R. H. Evans and W. W. Broenkow (1983), Phytoplankton pigment concentrations in the Middle Adantic Bight: comparison of ship determinations and CZCS estimates, Appl. Opt., 22, 20-36.

Gordon, L. I., L. A. Codispoti, J. C.Jennings, F.J. Millero, J. M. Morrison and C. Sweeney (2000), Seasonal evolution of hydrographic properties in the Ross Sea, Antarctica, 1996-1997, Deep Sea Res., Part II, 47(15-16), 3095-3117, doi:10.1016/S0967-0645(00)00060-6.

Greischar L. L., J. D. Robertson, C. R. Bendey and J. W. Clough (1981), Sea-bottom topography and crustal structure below the Ross Ice Shelf, Antarctica, in Antarctic Geology and Geophysics, edited by C. Craddock, pp. 1083-1090, University of Wisconsin Press, Madison WI.

Grigg, S. B. and N. J. Holbrook (2001), The impact of polynyas on the stability of the thermohaline circulation as simulated in a coupled ocean-atmosphere-sea ice box model, Geophys. Res. Lett., 28, 767-770.

Page 138: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Groisman, P. Y., R. W. Knight, D. R. Easterling, T. R. Karl, G. C. Hegerl and V. A. N. Razuvaev (2005), Trends in intense precipitation in the climate record, j.Clim., 18(9), 1326-50.

Haine, T. W. N. and K. J. Richards (1995), The influence of the seasonal mixed layer on oceanic uptake of CFCs,/. Geophys. Res., 100,10727-10744.

Heath, R. A. (1981), Oceanic fronts around southern New Zealand, Deep Sea Res., Part A, 28, 547-560.

Hellmer, H. H. (2004), Impact of Antarctic ice shelf basal melting on sea ice and deep ocean properties, Geophys. Res. Lett., 5/(10), GL019506, doi:doi:10.1029/2004GL019506.

Hibler, W. D. (1979), A dynamic thermodynamic sea ice model,/. Phys. Oceanogr., 9, 815-846.

Hofmann E. E. and J. M. Klinck (1998), Hydrography and Circulation of the Antarctic Continental Shelf: 150 E Eastward to the Greenwich Meridian, in The Global Coastal Ocean, Regional Studies and Synthesis, The Sea: Ideas and Observations in the Study of the Seas, vol. 11, edited by A. R. Robinson and K. H. Brink, pp. 997-1042, John Wiley & Sons, Inc., New York.

Holland, D. M. (1998), On the parameterization of basal heat flux for sea-ice modeling, Geophysica, 34,1-21.

Holland, D. M. (2001), Explaining the Weddell Polynya-a large ocean eddy shed at Maud Rise, Science, 292(5522), 1697-1700, doi:l0.1126/science. 1059322.

Holland, D. M., S. S. Jacobs and A. Jenkins (2003), Modelling the ocean circulation beneath the Ross Ice Shelf, Antarct. Sci., /5(1), 13-23, doi:10.1017/S0954102003001019.

Holland, D. M. and A. Jenkins (1999), Modeling thermodynamic ice-ocean interactions at the base of an ice shelf, / . Phys. Oceanogr., 29,1787-1800.

Holland, D. M. and A. Jenkins (2001), Adaptation of an isopycnic coordinate ocean model for the study of circulation beneath ice shelves, Mon. Weather Rev., 129, 1905-1927.

Holland, D. M., L. A. Mysak, D. K. Manak and J. M. Oberhuber (1993), Sensitivity study of a dynamic thermodynamic sea ice model,/. Geophys. Res., 98, 2561-2586.

Hunke, E. C. and S. F. Ackley (2001), A numerical investigation of the 1997-1998 Ronne polynya,/. Geophys. Res., 106,22373-22382.

IPCC, 2001: Climate Change 2001: The Science of"Climate Change. Contribution of the Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change,

Page 139: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Houghton, J.T., Y. Ding, D.J. Griggs, M. Noguer, P.J. van der Linden, X. Dai, K. Maskell, C.A. Johnson (eds.). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 881 pp.

IPCC, 2007: Climate Change 2007: The Physical Science Basis. Contribution of the Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 996 pp.

Jacobs, S. S., A. F. Amos and P. M. Bruchhausen (1970), Ross Sea oceanogrphy and Antarctic Bottom Water formation, Deep-Sea Res., 17, 935-962.

Jacobs, S. S. and J. C. Comiso (1989), Sea Ice and oceanic processes on the Ross Sea continental shelf,/. Geophys. Res., 94, 18195-18211.

Jacobs S. S., R. Fairbanks and Y. Horibe (1985), Origin and evolution of water masses near the Antarctic continental margin: Evidence from H2

1 80/H21 60 ratios in

seawater, in Oceanology of the Antarctic Continental Shelf, Antarc. Res. Sen, vol. 43, edited by S. Jacobs, pp. 59-85, AGU, Washington, D. C.

Jacobs S. S. and C. F. Giulivi (1999), Thermohaline Data and Ocean Circulation on the Ross Sea Continental Shelf, in Oceanography of the Ross Sea Antarctica, edited by G. Spezie and G. M. R. Manzella, pp. 3-16,.

Jacobs, S. S., A. L. Gordon and J. L. Ardai Jr. (1979), Circulation and Melting Beneath the Ross Ice Shelf, Science, 203, 439-443.

Jacobs, S. S., H. H. Hellmer, C. S. M. Doake, A. Jenkins and R. M. FroHch (1992), Melting of Ice Shelves and the Mass Balance of Antarctica,/. Glaciol, 38(\2>Q), 375-387.

Jacobs, S. S., H. H. Hellmer and A. Jenkins (1996), Antarctic ice sheet melting in the Southeast Pacific, Geophys. Res. Lett., 23(9), 957-960.

Jeffries, M. O. and U. Adolphs (1997), Early winter ice and snow thickness distribution, ice structure and development of the western Ross Sea pack ice between the ice edge and the Ross Ice Shelf, Antarct. Sci., 9,188-200.

Jenkins, A. and D. Holland (2007), Melting of floating ice and sea level rise, Geophys. Res. Lett., 34, L16609, doi:10.1029/2007GL030784.

Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S. Saha, G. White, J. Woollen, Y. Zhu, M. Chelliah, W. Ebisuzaki, W. Higgins, J. Janowiak, K. C. Mo, C. Ropelewski, J. Wang, A. Leetmaa, R. Reynolds, R. Jenne and D.Joseph (1996), The NCEP/NCAR 40-year reanalysis project, Bull. Am. Meteorol. Soc, 77,437-471.

Page 140: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Keys, H.J. R., S. S.Jacobs and L. W. Brigham (1998), Continued northward expansion of the Ross Ice Shelf, Antarctica, Annals of Glaciology, Vol27, 1998, 27, 93-98.

Killops S. K. and V. Killops (2005), Introduction to Organic Geochemistry, 393 pp., Blackwell Publishing, Maiden, MA.

Koentopp, M., O. Eisen, C. Kottmeier, L. Padman and P. Lemke (2005), Influence of tides on sea ice in the Weddell Sea: Investigations with a high-resolution dynamic-thermodynamic sea ice model,/. Geophys. Res., 110, C02014, doi: 10.1029/2004JC002405.

Lamy, F., J. Kaiser, H. W. Arz, D. Hebbeln, U. Ninnemann, O. Timm, A. Timmermann and J. R. Toggweiler (2007), Modulation of the bipolar seesaw in the Southeast Pacific during Termination 1, Earth Planet.Sci.Lett., 259(3-4), 400-13.

Lauriano, G., M. Vacchi, D. Ainley and G. Ballard (2007), Observations of top predators foraging on fish in the pack ice of the southern Ross Sea, Antarct.Sci., 19(4), 439-440.

Le Quere, C , C. Rodenbeck, E. T. Buitenhuis, T. J. Conway, R. Langenfelds, A. Gomez, C. Labuschagne, M. Ramonet, T. Nakazawa, N. Metzl, N. Gillett and M. Heimann (2007), Saturation of the Southern Ocean C02 sink due to recent climate change, Science (Washington D C), 316(5832), 1735-1738.

Legates, D. R. and C. J. Willmott (1990), Mean Seasonal and Spatial Variability in Gauge-Corrected, Global Precipitation, Int. J. Clim., 10,111-127.

Lindsay, R. W. and J. Zhang (2005), The thinning of Arctic sea ice, 1988-2003: have we passed a tipping point? J.Clim., 18(22), 4879-94.

Liss, P. S. and P. G. Slater (1974), Flux of gases across the air-sea interface, Nature, 247(5438), 181-4.

Lo Monaco, C, C. Goyet, N. Metzl, A. Poisson and F. Touratier (2005), Distribution and inventory of anthropogenic C02 in the Southern Ocean: Comparison of three data-based methods, Journal of Geophysical Research-Oceans, 110, C09S02.

Locarnini R. A. (1994), Water Masses and Circulation in the Ross Gyre and Environs, Ph.D. thesis, 87pp., Texas A7M University, Texas, December.

Lovenduski, N. S. and N. Gruber (2005), Impact of the Southern Annular Mode on Southern Ocean circulation and biology, Geophys.Res.Lett., 32(11), 4.

Lucarini, V. and G. L. Russell (2002), Comparison of mean climate trends in the Northern Hemisphere between National Centers for Environmental Prediction and two atmosphere-ocean model forced runs, journal of Geophysical Research, 107, ACL7-1.

Lucas, L. V. and J. E. Cloern (2002), Effects of tidal shallowing and deepening on phytoplankton production dynamics: A modeling study, Estuaries, 25, 497-507.

Page 141: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

MacAyeal, D. R. (1984), Numerical simulations of the Ross Sea tides,/. Geophys. Res., 89, 607-615.

Maqueda, M. A. M., A. J. Willmott and N. R. T. Biggs (2004), Polynya dynamics: A review of observations and modeling, Rev. Geophys., 42, RG1004, doi:10.1029/2002RG000116.

Marshall, G. J. (2003), Trends in the southern annular mode from observations and teanalyses,/.GEw., 16(24), 4134-43.

Marshall, G.J. (2007), Half-century seasonal relationships between the Southern Annular Mode and Antarctic temperatures, Int.J.Climatol, 27(3), 373-83.

Maslowski, W., D. C. Marble, W. Walczowski and A. J. Semtner (2001), On large-scale shifts in the Arctic Ocean and sea-ice conditions during 1979-98, Annals of'Glaciology, 33, 545-50.

Massom, R. A., P. T. Harris, K. J. Michael and M. J. Potter (1998), The distribution and formative processes of latent-heat polynyas in East Antarctica, Ann. GlacioL, 27, 420-426.

Massom, R. A., K. L. Hill, V. I. Lytle, A. P. Worby, M. Paget and I. Allison (2001), Effects of regional fast-ice and iceberg distributions on the behaviour of the Mertz Glacier polynya, East Antarctica, Ann. GlacioL, 33, 391-398.

McClain, C. R., K. R. Arrigo, K. S. Tai and D. Turk (1996), Observations and simulations of physical and biological processes at ocean weather station P, 1951-1980,/. Geophys. Res., 101(02), 3697-713.

McClatchie, S., R. B. Millar, F. Webster, P.J. Lester, R. Hurst and N. Bagley (1997), Demersal fish community diversity off New Zealand: Is it related to depth, latitude and regional surface phytoplankton? Deep Sea Res., Part I, 44, 647-667.

McPhee, M. G. (1980), An analysis of pack ice drift in summer, Sea Ice Processes and Models. Proceedings of the Arctic Ice Dynamics Joint Experiment International Commission on Snow and Ice Symposium, Univ. Washington Press, Seattle, Wash., 6-9 September.

McPhee, M. G. (1983), Turbulent heat and momentum transfer in the oceanic boundary layer under melting pack ice, / . Geophys. Res., 88, 2827-2835.

Mellor, G. L. and T. Yamada (1982), Development of a turbulence closure model for geophysical fluid problems, Rev. Geophys. Space Phys., 20, 851-875.

Michel, R. L., T. W. Linick and P. M. Williams (1979), Tritium and Carbon-14 Distributions in Seawater from Under the Ross Ice Shelf Project Ice Hole, Science, 203, AAS-AA6.

Page 142: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Monod J. (1942), Recherces sur la croissance des cultures bacteriennes, Herman et Cie, Paris.

Moore, M. I. and R. C. Murdoch (1993), Physical and biological observations of coastal squirts under nonupwelling conditions,/. Geophys. Res., 98, 20043-20061.

Morel, A. (1988), Optical modeling of the upper ocean in relation to its biogenous matter content (case I waters),/. Geophys. Res., 93(C9), 10749-68.

Ohshima, K. I. and S. Nihashi (2005), A simplified ice-ocean coupled model for the Antarctic ice melt season,/. Phys. Oceanogr., 35(2), 188-201, doi:10.1175/JPO-2675.1.

Olbers, D.J., Gouretzki, V., Seiss, G. & Schroter, J. 1992. The hydrographic atlas of the Southern Ocean. Bremerhaven: Alfred Wegener Institute, http://www.awi-bremerhaven.de/Atlas/SO

Olson, R. J., H. M. Sosik, A. M. Chekalyuk and A. Shalapyonok (2000), Effects of iron enrichment on phytoplankton in the Southern Ocean during late summer: active fluorescence and flow cytometric analyses, Deep Sea Res., Part II, 47, 3181-3200.

Overland, J. E. and K. L. Davidson (1992), Geostrophic drag coefficient over sea ice, Dynam. Met. and Oceanography, 44A, 54-66.

Overland, J. E., H. O. Mofjeld and C. H. Pease (1984), Wind-Driven Ice Drift in a Shallow Sea,/. Geophys. Res., 89, 6525-6531.

Padman, L., S. Erofeeva and I. Joughin (2003), Tides of the Ross Sea and Ross Ice Shelf cavity, Antarct. Set., 15(1), 1-10, doi: 10.1017/S0954102002.

Pease, C. H. (1987), The size of wind-driven coastal polynyas,/. Geophys. Res., 92, 7049-7059.

Picco P., L. Amici, R. Meloni, L. Langone and M. Ravaioli (1999), Temporal variability of currents in the Ross Sea (Antarctica), in Oceanography of the Ross sea Antarctica, edited by G. Spezie and G. M. R. Manzella, pp. 103-117, Springer, New York.

Polito, M., S. D. Emslie and W. Walker (2002), A 1000-year record of Adelie penguin diets in the southern Ross Sea, Antarct. Sci, 14(A), 327-332, doi:10.1017/S0954102002000184.

Polyakov, I. and S. Martin (2000), Interaction of the Okhotsk Sea diurnal tides with the KashevarovBankpolynya,/. Geophys. Res., 105, 3281-3294.

Prezelin, B. B., E. E. Hofmann, C. Mengelt and J. M. Klinck (2000), The linkage between Upper Circumpolar Deep Water (UCDW) and phytoplankton assemblages on the west Antarctic Peninsula continental shelf,/. Mar. Res., 58, 165-202.

129

Page 143: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Reddy, T. E. and K. R. Arrigo (2006), Constraints on the extent of the Ross Sea phytoplankton bloom,/. Geophys. Res., 111, C07005, doi:10.1029/2005JC003339.

Reddy, T. E. and K. R. Arrigo (in prep), Ross Sea Primary Production and Sea Ice under Future Climatic Change Scenarios, / . Geophys. Res.

Reddy, T. E., K. R. Arrigo and D. M. Holland (2007), The Role of Thermal and Mechanical Processes in the Formation of the Ross Sea Summer Polynya,/. Geophys. Res., doi:10.1029/2006JC003874.

Reddy, T. E., D. M. Holland and K. R. Arrigo (submitted), Ross Ice Shelf Cavity Circulation, Residence Time, and Melting: Results from a Model of Oceanic Chlorofluorocarbons, Umnol. Oceanogr.

Richardson, A. J. and D. S. Schoeman (2004), Climate impact on plankton ecosystems in the Northeast Atlantic, Science, 305(5690), 1609-12.

Roberts, A., I. Allison and V. I. Lytle (2001), Sensible- and latent-heat-flux estimates over the Mertz Glacier polynya, East Antarctica, from in-flight measurements, Ann. Glaciol., 33, 377-384.

Robertson R. A., L. Padman and G. D. Egbert (1998), Tides in the Weddell Sea, in Ocean, Ice, and Atmosphere: Interactions at the Antarctic Continental, Antarct. Res. Sen, vol. 75, edited by S. S.Jacobs and R. F. Weiss, pp. 341-369, AGU, Washington, D.C.

Roman, M., X. Zhang, C. McGilliard and W. Boicourt (2005), Seasonal and annual variability in the spatial patterns of plankton biomass in Chesapeake Bay, Umnol. Oceanogr., 50, 480-492.

Russell, J. L., K. W. Dixon, A. Gnanadesikan, R. J. Stouffer and J. R. Toggweiler (2006), The Southern Hemisphere westerlies in a warming world: propping open the door to the deep oce&n,J.Clim., 19(24), 6382-90.

Sarmiento, J. L., T. M. C. Hughes, R. J. Stouffer and S. Manabe (1998), Simulated response of the ocean carbon cycle to anthropogenic climate warming, Nature, 393(6682), 245-9.

Sarmiento, J. L. and C. LeQuere (1996), Oceanic carbon dioxide uptake in a model of century-scale global warming, Science, 274(5291), 1346-50.

SCAR. 1993. Antarctic Digital Database. Cambridge: Scientific Committee on Antarctic Research, http://www.nercbas.ac.uk/public/magic/add.html

Semtner, A. J. and R. M. Chervin (1992), Ocean general circulation from a global eddy-resolving model,/. Geophys. Res., 97, 5493-5550.

Shillington, F. A., W. T. Peterson, L. Hutchings, T. A. Probyn, H. N. Waldron and J. J. Agenbag (1990), A cool upwelling filament off Namibia, southwest Africa:

130

Page 144: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

preliminary measurements of physical and biological features, Deep Sea Res., Part A, 37,1753-1772.

Shimada, K., T. Kamoshida, M. Itoh, S. Nishino, E. Carmack, F. McLaughlin, S. Zimmermann and A. Proshutinsky (2006), Pacific Ocean inflow: Influence on catastrophic reduction of sea ice cover in the Arctic Ocean, Geophys.Res.Lett., 33(8), L08605.

Shindell, D. T., R. L. Miller, G. A. Schmidt and L. Pandolfo (1999), Simulation of recent northern winter climate trends by greenhouse-gas forcing, Nature, 399(6735), 452-455.

Shindell, D. T. and G. A. Schmidt (2004), Southern Hemisphere climate response to ozone changes and greenhouse gas increases, Geophys.Res.Lett., i/(18), LI8209-4.

Simmonds, I. and K. Keay (2000), Variability of Southern Hemisphere extratropical cyclone behavior, 1958-97, J.Clim., 13(3), 550-61.

Smethie, W. M. and S. S. Jacobs (2005), Circulation and melting under the Ross Ice Shelf: estimates from evolving CFC, salinity and temperature fields in the Ross Sea, Deep Sea Res. I, 52(6), 959-978, doi:10.1016/j.dsr.2004.11.016.

Smith, S. D., R. D. Muench and C. H. Pease (1990), Polynyas and leads: an overview of physical processes and environment,/. Geophys. Res., 95, 9461-9479.

Smith, W. O. and L. I. Gordon (1997), Hyperproductivity of the Ross Sea (Antarctica) polynya during austral spring, Geophys. Res. Lett., 24, 233-236.

Steele, M. and T. Boyd (1998), Retreat of the cold halocline layer in the Arctic Ocean, Journal of Geophysical Research, 103, 10419-35.

Stewart R. H. (2004), Introduction to Physical Oceanography, 344 pp., Texas A&M University, College Station.

Stouffer, R. J., S. Manabe, K. Bryan and K. Aagaard (1989), Interhemispheric asymmetry in climate response to a gradual increase of atmospheric C0 2 , Nature, 342(6250), 660-662.

Sweeney, C , W. O. Smith, B. Hales, R. R. Bidigare, C. A. Carlson, L. A. Codispoti, L. I. Gordon, D. A. Hansell, F. J. Millero, M. O. Park and T. Takahashi (2000), Nutrient and carbon removal ratios and fluxes in the Ross Sea, Antarctica, Deep Sea Res., Part 11,47,3395-3421.

Trumbore, S. E., S. S.Jacobs and W. M. Smethie Jr (1991), Chlorofluorocarbon evidence for rapid ventilation of the Ross Sea, Deep-Sea Research, Part A (Oceanographic Research Papers), 38(1 A), 845-70.

Page 145: OCEANOGRAPHY OF THE ROSS SEA: OCEAN ...ocean.stanford.edu/people/dissertations/reddy...LIST OF FIGURES x CHAPTER 1: Introduction 1 CHAPTER 2: Constraints on the Extent of the Ross

Van Woert, M. L. (1999), Wintertime dynamics of the Terra Nova Bay polynya,/. Geophys. Res., 104, 7753-7769.

Van Woert M. L., E. S. Jahnson, L. Langone, D. L. Worthen, A. Monaghan, D. H. Bromwich, R. Meloni and R. B. Dunbar (2003), The Ross Sea Circulation During the 1990s, in Biogeochemistry of the Ross Sea, Antarct. Res. Ser., vol. 78, edited by G. R. DiTullio and R. B. Dunbar, pp. 5-34, AGU, Washington, D.C.

Van Woert, M. L., W. N. Meier, C. Z. Zou, A. Archer, A. Pellegrini, P. Grigioni and C. Bertola (2001), Satellite observations of upper-ocean currents in Terra Nova Bay, Antarctica, Ann. Glaciol, 33, 407-412.

Walker, S. J., R. F. Weiss and P. K. Salameh (2000), Reconstructed histories of the annual mean atmospheric mole fractions for the halocarbons CFC-11, CFC-12, CFC-113, and carbon tetrachloride, Journal of Geophysical Research, 105(C6), 14285-96.

Wanninkhof, R. (1992), Relationship between wind speed and gas exchange over the ocean, Journal of'Geophysical'Research, 97(C5), 7373-7382.

Warner, M.J. and R. F. Weiss (1985), Solubilities of chlorofluorocarbons 11 and 12 in water and seawater, Deep Sea Res., 32(12A), 1485-1497.

Winton, M. (2006), Does the Arctic sea ice have a tipping point? Geophjs.Res.Lett., 33(23), 1-5.

Worthen D. L. and K. R. Arrigo (2003), A coupled ocean-ecosystem model of the Ross Sea: 1. Interannual variability of primary production and phytoplankton community structure, in Biogeochemistry of the Ross Sea, Antarct. Res. Ser., vol. 78, edited by G. R. DiTullio and R. B. Dunbar, pp. 93-105, AGU, Washington, D.C.

Wu, X. G., W. F. Budd and I. Allison (2003), Modelling the impacts of persistent Antarctic polynyas with an atmosphere-sea-ice general circulation model, Deep Sea Res., Part II, 30(8-9), 1357-1372, doi:10.1016/S0967-0645(03)00072-9.

Wu, X. G., I. Simmonds and W. F. Budd (1997), Modeling of Antarctic sea ice in a general circulation model,/. Clim., 10, 593-609.

Zheng, M., W.J. De Bruyn and E. S. Saltzman (1998), Measurements of the diffusion coefficients of CFC-11 and CFC-12 in pure water and seawater,/. Geophys. Res., 103(d), 1375-1379.

Zwally H. J., J. C. Comiso and A. L. Gordon (1985), Antarctic offshore leads and polynyas and oceanographic effects, in Oceanology of Antarctic Continental Shelf Antarct. Res. Ser., vol. 43, edited by S. S.Jacobs, pp. 203-226, AGU, Washington, D.C.

132