Quick viewing(Text Mode)

Spacial Variability in Plankton Size Structure and Community Composition Along Biogeochemical Gradients in the Pacific Ocean

Spacial Variability in Plankton Size Structure and Community Composition Along Biogeochemical Gradients in the Pacific Ocean

UNIVERSITY OF HAWAII LIBRARY

SPACIAL VARIABILITY IN SIZE STRUCTURE AND COMPOSITION ALONG BIOGEOCHEMICAL GRADIENTS IN THE PACIFIC

A TIIESIS SUBMIITED TO THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI'! IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF SCIENCE

IN

OCEANOGRAPHY

AUGUST 2007

By Tara M. Clemente

Thesis Committee:

David M. Karl, Chairperson Matthew J. Church Karin M. Bjorkman We certify that we have read this thesis and that, in our opinion, it is satisfactory in scope and quality as a thesis for the degree of Master of Science in Oceanography. ACKNOWLEDGEMENTS

I would like to thank my committee members for their time, encouragement, support and advice. My primary advisor, Dave Karl, who has challenged me to think critically and guided me throughout this study; Matt Church, for always leaving his door open and for giving me insight throughout this project; and Karin BjOrkman, who has been instrumental throughout this entire project especially in both the field and lab. I would like to thank the members of the HOT team for all their support and for their friendships.

I'd like to thank the officers and crew of the RfV Kilo Moana for providing a safe and comfortable work place. Thanks to all the BEACH-BASH participants for their hard work and assistance with sample collection and data analysis, in particular Chief Scientist

Karin BjOrkman, Eric Grabowski and Dan Sadler. For help in analysis I wish to acknowledge Bob Bidigare and Stephanie Christensen for HPLC, Karin BjOrkman, Susan

Curless and Claire Mahaffey for dissolved nutrients and Karen Selph for .

I'd like to thank my friends who stood by me throughout these last couple of years and gave me countless hours of advice, support and ice cream. I'd especially like to thank my family, my mom and dad for always supporting my goals no matter how strange they may have been at times; and for my sisters Jamie and Tracey, who were always available for those late night phone calls. I wish to give a special thanks to David Nichols whose support, friendship, and ability to make me laugh throughout these many years has been incredible. Lastly, I acknowledge the National Science Foundation, the Gordon and

Betty Moore Foundation, and the Hawaii Ocean Time-series for providing financial support.

iii ABSTRACT

The size structure of upper ocean plankton assemblages appears to play an important role in determining the efficiency of the biological carbon pump and the resulting magnitude of biologically-mediated carbon export to the . To evaluate spatial variability in the size structure and community composition of plankton assemblages in the upper ocean measurements of size-fractionated a (Chi a), phycoerythrin and adenosine 5' -triphosphate (A1P) concentrations along with abundances and taxonomic pigment biomarkers were assessed during the BEACH-BASH transect cruise from American Samoa to Honolulu, Hawaii in March 2005. Sampling for this study spanned four different biogeochemical provinces including the oligotrophic South Pacific

Subtropical Gyre (SPSG), the relatively nutrient-enriched equatorial upwelling provinces, the Pacific Equatorial Divergence (PEQD) and the North Pacific Equatorial

Countercurrent (PNEC), and the oligotrophic North Pacific Tropical Gyre (NPTG).

Overall, nutrient concentrations were highest in the PEQD province, coinciding with the strong westerly South Equatorial Current (SEC) and the region of equatorial upwelling, and lowest in the SPSG and NPTG. Chi a and phycoerythrin concentrations throughout the transect were overwhelmingly dominated by picoplankton (0.2-2 f.UD.), accounting for

45 - 60 % of the total Chi a and 80 - 95 % of the total phycoerythrin throughout the transect. Flow cytometry measurements indicate that the picoplankton community composition was dominated by of two genera; Prochlorococcus spp. and

Synechococcus spp. and these groups accounted for most of the Chi a throughout the transect. Picoeukaryotes also contributed to picoplankton abundances; however

iv despite the likelihood of many species being present they remain largely unidentified.

Total microbial biomass estimates via particulate ATP (P-ATP) measurements showed high variability with the occasional occurrence of larger heterotrophic organisms

(>20 J.Lffi and 2-10 J.Lffi) which are not accounted for in Chi a biomass estimates.

However, their distribution is sure to play an important role in regulating size structure and food web dynamics. Measurements of taxonomic pigment markers by HPLC confirmed the overwhelming dominance of prochlorophytes and cyanobacteria throughout the transect, however a transition in community composition in the PEQD province was observed. This transition was attributed to the increase in taxonomic pigment markers of chromophyte (i.e. prymnesiophytes, pelagophytes and ).

v TABLE OF CONTENTS

Section

Acknowledgements ...... ii

Abstract ...... iv

List of Tables ...... viii

List of Figures ...... ix

Chapter I. Introduction ...... 1

Chapter 2. Spatial variability in plankton size structure and community composition

along biogeochemical gradients in the Pacific Ocean ...... •...... 5

2.1 Introduction ...... 5

2.1.1 Plankton size structure and community composition ...... 5

2.1.2 Biogeochemical provinces ...... 8

2.1.3 Equatorial Pacific Ocean ...... 11

2.1.4 Size-fractionation ...... 16

2.1.5 Objectives of this study ...... 17

2.2 Methods ...... 18

2.2.1 BEACH-BASH data ...... 18

2.2.2 Sample collection and analysis ...... 18

2.2.2.1 Nutrient analysis ...... 18

2.2.2.2 Chlorophyll a ...... 21

2.2.2.3 Phycoerythrin ...... •...... 22

2.2.2.4 Adenosine 5'-triphosphate (ATP) ...... 23

2.2.2.5 Flow cytometry ...... 24

vi 2.2.2.6 HPLC pigments ...... 25

2.3 Results ...... 28

2.3.1 Habitat characteristics ...... 28

2.3.1.1 Biogeochemical provinces ...... 28

2.3.1.2 Latitudinal distribution of temperature, salinity

and Chl a ...... 31

2.3.2 Nutrient distributions ...... 33

2.3.2.1 Latitudinal distribution of dissolved inorganic nutrients.. 33

2.3.2.2 Dissolved inorganic nutrient ratios ...... •...... 35

2.3.3 Plankton size structure ••••••••••••••••••.•.•••••••••••••••••••••.•.•.••.• 37

2.3.3.1 Chlorophyll a ...... 37

2.3.3.2 Phycoerythrin ...•...... •..42

2.3.3.3 Adenosine 5'-triphosphate (ATP) ...... 44

2.3.4 Picoplankton community structure and taxonomic distribution ...... 47

2.3.4.1 Flow cytometry ...... •...... 47

2.3.4.2 HPLC pigments ...... 51

2.4 Discussion ...... 59

2.4.1 Nutrient dynamics ...... 59

2.4.2 size structure ...... 62

2.4.3 Community structure and taxonomic composition ...... 67

2.4.4 Implications for the ...... •...... 75

Chapter 3. Conclusions ...... 77

References ...... 79

vii LIST OF TABLES

1. Detection limits, precision estimates and observed ranges .•...... 20

2. Abbreviations and taxonomic affinities of photosynthetic pigments ...... •...... 26

viii LIST OF FIGURES

Figures

1. Conceptual view of major controls on recycled vs. export production ...... • 6

2. Classification of planktonic organisms based on size ...... 7

3. BEACH-BASH transect cruise track ...... 9

4. Biogeochemical provinces sampled during BEACH-BASH ...... 10

5. ADCP data showing current velocities ...... 12

6. Sea surface temperature, wind and anomalies from TAD arrays ...... 15

7. Depth profiles of temperature, chlorophyll a and soluble reactive phosphorus in each of the biogeochemical provinces ...... 29

8. Latitudinal distribution of sea surface temperature, chlorophyll a and salinity ...... 32

9. Latitudinal distribution of surface nutrients nitrate + nitrite, soluble reactive phosphorus, and silicic acid ...... 34

10. Latitudinal distribution of surface nutrient ratios ...... 36

11. Latitudinal distribution of surface size-fractionated chlorophyll a ...... 38

12. Chlorophyll a regression analysis of GFIF vs. summed PC filters ...... 40

13. Correlations between chlorophyll a size fractions and nutrients -...... 41

14. Latitudinal distribution of surface size-fractionated phycoerythrin ...... 43

15. Latitudinal distribution of surface total particulate A TP (P-A TP) ...... 45

16. Latitudinal distribution of surface size-fractionated P-ATP ...... 46

17. Latitudinal distribution of picoplankton abundances ...... 48

18. Correlations between surface abundance of plankton groups and nutrients ...... 49

19. Latitudinal distribution ofHPLC TChl a, TChl band TChl c ...... 53

ix Figures

20. Chlorophyll a regression analysis ofTD-700 vs. HPLC ...... 54

21. Latitudinal distribution ofTChl a constituents MVChl a, DVChl a and MVChld a ...... 55

22. CyanoDP and ChromoDP as a proportion of total diagnostic pigments ...... 57

23. Temporal distribution of Prochlorococcus spp. abundance at Station ALOHA between 1991-2005 and the Southern Oscillation Index (SOl) ...... 70

x CHAPTER 1. INTRODUcnON

The production of organic matter in the sea is sustained by the continuous supply of the bioessential elements including carbon (C), nitrogen (N), phosphorus (P), and silica (Si) as well as minor nutrients such as iron (Fe) and zinc (Zn). Redfield et al. (1963) summarized the average bulk elemental stoichiometry of plankton particulate organic matter as a ratio ofl06C: 16N: IP; however, the chemical composition of living organisms varies as a function of growth rate, energy availability, ambient nutrient concentrations and nutrient concentration ratios (Sakshaug and Holm-Hansen, 1977;

Rhee, 1978; Laws and Bannister, 1980; Tett et af., 1985). In large areas of the world's , low concentrations ofbioavailable nutrients control production and plankton biomass. Liebig's Law of the Minimum states that ecosystem productivity is limited by the nutrient which is exhausted first relative to its demand (Liebig, 1840). The supply of nutrients to the productive surface layers of the open ocean and marine inhabitants can occur by large-scale processes such as eddy diffusion, upwelling, atmospheric deposition, land runoff or by loca1ized process such as the flux of remineralized nutrients within the (Azam et al., 1983; Dugdale and

Goering, 1967).

Microorganisms in pelagic are generally classified on the basis of size, nutritional and physiological characteristics or phylogeny (Karl, 1999). The size of a cell can have many consequences on the fitness of an organism in a habitat and on its ability to use resources like light and nutrients toward a competitive advantage (Raven, 1986).

1 Sheldon el a1. (1972) were the first to supply an image of size distributions of plankton across a wide expanse of the world ocean. Sheldon put forth the Linear Biomass

Hypothesis, which states that the amount of biomass in equally-sized logarithmic size classes remains constant through the food web from to whales (Sheldon et al.,

1972). This hypothesis, promoted before the was envisioned, provided a simple image of community structure, and provided a means for predicting fish stocks given the biomass of plankton (Sheldon et a!., 1977). Since the Sheldon et al. (1972) study, which used a coulter counter to investigate size structure, improved methods such as flow cytometry have been developed which has the ability to characterize small cells rapidly to generate size structure data (platt and Denman, 1977; Blanco et a1., 1994;

Vidondo et al., 1997). This enabled the discovery of two genera of picoplankton

Prochlorococcus spp. and Synechococcus spp. which are now known to contribute significantly to overall and autotrophic biomass in oligotrophic marine environments (Chisholm, 1992). Prochlorococcus spp. was discovered about two decades ago in the Sargasso Sea by the use of on-board flow cytometry (Chisholm et al.,

1988; 1992). It was 1ater reported in the equatorial Pacific (Chavez et a1., 1991) and is now considered to be of global significance (partensky et al., 1999). Synechococcus spp. was first described in 1979 (Johnson and Sieburth, 1979; Waterbury et a1., 1979) and is now known to be an important component of the picoplankton community in temperate to tropical oceans.

Plankton cell sizes are typically characterized as picoplank:ton (0.2-2 1IJIl); nanoplankton

(2-20 1IJIl); microplankton (20-200 1IJIl); and mesoplankton (0.2-2 mm; Sieburth et al.,

2 1978}. Large cells, such as diatoms are observed to dominate on a biomass basis in high nutrient areas while small cells such as cyanobacteria tend to dominate oligotrophic areas of the world's oceans (Li et al., 1983; Hagstrllm et al., 1988). The size structure and biomass of upper ocean plankton assemblages can help provide an understanding of trophic structure, energy flow and export production within an ecosystem (Karl and

Dobbs, 1998). Legendre and Le Fevre (1989) suggest that export production is maximized iflarge phototrophs grow, aggregate and then sink and that export is minimized when phototrophic picoplankton fuel a complex microbial food web. Based on this conceptual framework, it appears that the size structure of plankton assemblages in the open ocean can play an important role in determining the efficiency of the biological carbon pump and the resulting magnitude of biologically-mediated carbon export to the deep sea.

There have been a large number of investigations on the spatial distribution and temporal variability of phytoplankton around the world's oceans. In the Atlantic Ocean the

Atlantic Meridional Transect (AMT) research program conducted several oceanographic cruises along a transect from 50' S to 50' N encompassing various biogeochemical provinces. Several of these studies focused on phytoplankton size structure and productivity (Marafl.6n et al., 2000; 2001; 2003; Fernandez et al., 2003; perez et al.,

2005; San Martin et al., 2006) as well as microbial abundance and taxonomic composition (Zubkov et al., 1998; 2000; Gibb et al., 2000; Barlow et al., 2002; 2004;

Poulton et al., 2006). In the North Atlantic, Li and Harrison (2001) partitioned bacteria and picophytoplankton into ecological provinces as specified by Longhurst (1998), and

3 the relationships of nutrient gradients to microbial community structure were investigated in Cavender-Bares et al. (2001).

In the Pacific Ocean no one study has encompassed as wide of a latitudinal range as that sampled in the Atlantic Ocean during the AMT program. However, several cruises have been conducted by the Joint Global Ocean Flux Study (JGOFS) in the equatorial Pacific

Ocean. The objectives of the JOOFS studies were to determine the fluxes of carbon and related elements and processes controlling those fluxes between the euphotic zone, the atmosphere, and the deep ocean. The equatorial JGOFS cruises include EqPac and

Surtropac (1992), FLUPAC and OLIPAC (1994), and EBENE (1996) and together encompass a latitudinal range between 20° S - 12° N. These JGOFS cruises along with the Iron Ex II (1995) cruise, helped capture microbial dynamics at various temporal and spatial scales of variability (Landry and Kirchman, 2002). In the North Pacific Ocean, the consistently low nutrient subtropical waters of the long term sampling Hawaii Ocean

Time-series (HOn site, Station ALOHA, have previously been used as a frame of reference for Variability in equatorial regions. To my knowledge the study presented in this thesis is the first in the equatorial Pacific to investigate the spatial dynamics of phytoplankton along biogeochemical gradients over a wide latitudinal range that includes the HOT site Station ALOHA.

4 CHAPTER 2. SPATIAL VARIABLITIY IN PLANKTON SIZE STRUCTURE

AND COMMUNITY COMPOSmON ALONG BIOGEOCHEMCIAL

GRADIENTS IN THE PACIFIC OCEAN

2.1 INTRODUCTION

2.1.1 Plankton size structure and community composition

The size structure and community composition of phytoplankton assemblages in the open ocean are important biological factors that regulate the functioning of pelagic food-webs and the export of organic carbon to the deep ocean (Azam et aZ., 1983; Tremblay and

Legendre, 1994). Large cells, such as diatoms are observed to dominate on a biomass basis in high nutrient areas, have rapid sinking rates and are believed to be large contributors to carbon export (Michaels and Silver, 1988), while small cells such as cyanobacteria tend to dominate oligotrophic areas of the world's oceans (Li et aZ., 1983;

Hagstr6m et aZ., 1988) and are believed to contribute relatively little to carbon export because of their small sizes, slow sinking rates and rapid utilization in the microbial loop

(Legendre and Le Fevre, 1989; Michaels and Silver, 1998; Figure 1). Phytoplankton may be classified based on cell size and include small picoplankton (0.2-2 f.Illl); medium-sized nanoplankton (2-20 f.Illl); large microplankton (20-200 f.Illl); and even larger mesoplankton (0.2-2 mm; Sieburth et al., 1978; Figure 2). This study took place in the

Pacific Ocean and focuses on three of the four previously stated size classes; picoplankton (e.g. prochlorophytes, Synechococcus spp., small ), nanoplankton

(e.g. prymnesiophytes, pelagophytes, small diatoms and ) and microplankton (e.g. diatoms and dinoflagellates). It also further classifies nanoplankton 5 IDIA TOMS I IPROCHLOROCOCCUS I

MICROBIAL LOOP

IEXPORT FLUX I

Adapted fromiLesn1ruud Le.FiMe (1989)

Figure I. Conceptual view of major controls on recycled vs. export production in the sea. Image courtesy of David M. Karl.

6 Plankton Size categories (pm) 2000 0en w :E

200

-&J ~ ~ @~ f1 ~j1 ~ ~ 20 e wv- ~~ CI"V- ~ ,. ~~ ~ ~ ~ oJ'- --~-- 2

• • ~ • ~ • • .. ~ • • . . • • ~ • • • . . .. • .. • ... • • .. . . • • 0.2

Figure 2. Classification of planktonic organisms by size based on Sieburth el al. (1978) Limnology and Oceanography 23 : 1256-1 263 taken from Karl ( 1999). This study focused on three oftbe four size classes shown above; picoplankton (0 .2-2 flm) (e.g. prochlorophytes, Synechococcus spp., and small eukaryotes), nanoplankton (2-20 flm) (e.g. prymnesiophytes, pelagophytes, small diatoms and dinoflagellates) and microplankton (20-200 flm) (e .g. diatoms and dinoflagellates) and it further classifies nanoplankton into two size classes (2-10 flm and 10-20 flm).

7 into two size classes (2-10 pm and 10-20 pm). Nutrients and light aVailability along with predation and physical processes such as upwelling are factors that can influence the composition and dynamics of the phytop1ankton community. These factors all vary in time and space and can lead to variability in phytoplankton community composition and growth rates. Typically, the upper ocean is considered to be nutrient limited and well illuminated and the deep ocean is considered to be nutrient rich and light limited.

2.1.2 Biogeochemical provinces

To compare data over extensive spatial scales, stations sampled during the BEACH­

BASH (Biogeochemical and Ecological Analysis of Complex Habitats-Between

American Samoa and Hawaii) transect cruise (Figure 3) were grouped according to the biogeochemical provinces proposed by Longhurst (1998), which are based on global­ scale productivity inferred from remotely sensed ocean color. According to this approach, four different biogeochemical provinces were sampled (Figure 4). These included the South Pacific Subtropical Gyre (SPSG) province, Pacific Equatorial

Divergence (PEQD) province, North Pacific Equatorial Countercurrent (PNEC) province, and the North Pacific Tropical Gyre (NPTG) province. The SPSG province as defined by

Longhurst (1998) comprises the central and southern part of the subtropical gyre of the

South Pacific Ocean extending from 35° S to 5° S. It is bounded by the chlorophyll enhancement of the Subtropical Convergence Zone (SSTC) in the south and the southern edge of chlorophyll enhancement caused by equatorial divergence in the north. The

PEQD province corresponds with the zone of nitrate-replete water of the tropical Pacific and extends from 5° S across the equator to 5° N (Longhurst, 1998). The surface

8 American Samoa to Hawaii 0.3

0.25

::::: ~ 0.2 >- QI ~ "C Q. ::l ...0 ~ 0 -ns ~ ...J 0.15 () QI U ~ en::l ns 0.1 enQI

0.5

o -180 -175 -170 -165 -160 -155 -150 Longitude

Figure 3. MODIS satellite derived sea surface chlorophyll concentrations along the BEACH-BASH transect cruise track (April 29 - May 9, 2005).

9 PEQD

SPSG

,r"'l Longhurst 1998

Provinces Acron ym Location Stations sampled

South Pacific Subtropical Gyre SPSG 35°S - 50S N=5 - Pacific Equatorial Divergence PEQD 50S - 5°N N = 12 5°N - 100N N =4 North Pacific Equatorial Countercurrent PNEC - North Pacific Tropical Gvre NPTG 100N - 32°N N=6

Figure 4. Biogeochemical provinces sampled during the BEACH- BASH transect (shown in black) as defined by Longhurst (1998). SPSG (B lue), PEQD (White), PNEC (Yellow) and NPTG (Grey) 10 circulation in the PEQD is dominated by the flow of the South Equatorial Current (SEC) which moves westwards across the ocean (Figure 5). Within this circulation occurs a strong linear divergence and upwelling along the equator (Longhurst, 1998). This divergence is forced by the change in sign of the Coriolis force at the equator polewards in both hemispheres for a westward flow. The PNEC province lies between the northern and southern Doldrum salinity fronts at 5· N and 10· N (Longhurst, 1998). In this province the curvature of wind stress not only forces the flow of the North Equatorial

Countercurrent (NECC). but is also thought to enhance algal growth and therefore

Chlorophyll a (Chl a) concentrations (Longhurst, 1993; Figure 5). The NPTG province is very large and lies between the Subtropical Convergence at 32· N and the northern

Doldrum Front (Roden, 1975) at 10· N (Longhurst, 1998).

2.1.3 EqUllloriol Pacific Ocean

Approximately one-half of global primary production is supported by the photosynthetic activities of microscopic plankton in the world's oceans (Falkowski, 1994). Oceanic provinces far removed from land account for most of the primary production. The

Pacific Ocean is the largest water body in the world and can be divided into two parts according to the nutrient availability: the eutrophic region and the oligotrophic region. In the Pacific Ocean the eutrophic regions constitute the central equatorial upwelling zone and coastal areas due to the input of nutrients from land. However, the majority of the open ocean is made up of oligotrophic waters. It is therefore of interest to investigate the spatial dynamics of plankton size structure along biogeochemical gradients from the

11 Kilo Moana L 25 III to 35 m

CC

SEC

"'..

20 0 S~ ______~

1800 W 1600 W .r''''''''''ri ... '''''''''1, o 100

Figure 5. ADCP data showing the current velocities between the 25 m to 35 m depth layer during the BEACH-BASH transect cruise. South Equatorial Current (SEC), North Equatorial Countercurrent (NECC) and North Equatorial Current (NEC). Biogeochemical provinces are represented by colors; SPSG (Blue), PEQD (White), PNEC (Yellow) and NPTG (Grey). Image courtesy of Eric Firing and Jules Hurnmon.

12 South Pacific, across the equator to the North Pacific to determine the factors controlling their standing biomass.

The equatorial Pacific is extremely dynamic; it is characterized by episodic upwelling events, elevated chlorophyll concentrations and carbon fixation rates. The north and southeast tradewinds set up a shallow thermocline in the eastern and central Pacific and drive equatorial upwelling. The equatorial Pacific plays an important role in the global . Upwelling of cool water from below the shallow thermocline results in a large supply of macronutrients including carbon dioxide (C02) to the surface (Chavez and Barber, 1987; Feeley et al., 1987). It supplies annually about 1 gigaton (Ot) of carbon dioxide (C~) to the atmosphere and is the largest natural source of C02 to the atmosphere (Tans et al., 1990; Feeley et aI., 1987). At the same time, it also contributes a significant fraction of the ocean's new production (0.8 - 1.9 Ot C year -I; Chavez and

Barber, 1987). Even though concentrations of macronutrients are well above those that saturate phytoplankton uptake (Barber and Chavez, 1986), phytoplankton in the equatorial Pacific rarely if ever bloom and the depletion time of nitrate is long (Dugdale et al., 1992; Fieldler et al., 1991; McCarthy et al., 1996). The equatorial Pacific is also considered a region of high nutrient low chlorophyll (HNLC; Minas et al., 1986; Cullen,

1991). Several hypotheses have been examined to explain HNLC conditions in the equatorial Pacific. These included inhibition of phytoplankton growth by trace metals such as iron (Barber and Ryther, 1969; Martin, 1990), physical processes (Thomas,

1972), grazing control (Walsh, 1976; Pella et aI., 1990), regulation by silicate (Dugdale et al., 1995), and lack of bloom forming diatoms (Chavez, 1989).

13 The equatorial Pacific is also an area of dramatic variability in the physical environment.

Periodically the east-west slope in the equatorial Pacific thennocline is affected by El

Nifio or La Nifia (philander, 1990). During El Nifio the thennocline is depressed in the eastern Pacific and the supply of macronutrients to the surface waters is lower than normal, warm sea surface temperatures extend throughout the equatorial region due to relaxed tradewinds and reduced upwelling and the anomaly shows a large change in sea surface temperature that depicts warming in the eastern equatorial Pacific (Figure 6a).

During La Nifia the thennocline is even shallower than normal and the supply of macronutrients to surface waters is enhanced, cool sea surface temperatures extend throughout most of the basin due to increased winds and upwelling and the anomaly shows a large change in sea surface temperature that depicts cooling in the eastern equatorial Pacific (Figure 6b). Besides being influenced by the El Nifio-Southern

Oscillation (ENSO) cycle, the equatorial Pacific waters experience physical forcings that include Kelvin waves, Tropical instability waves (TIWs) and wind induced upwelling.

These physical phenomena can affect organism concentrations and their physiological characteristics on considerably different time and space scales making it important to employ appropriate sampling strategies (Dickey, 1991).

The equatorial Pacific is frequently divided into three different biogeochemical provinces

(Longhurst, 1998). Two of which I have mentioned previously, the PNEC and PEQD provinces which are located in the eastern and central equatorial Pacific and a third known as the Western Pacific Warm Pool (WARM) province which is located in the western Pacific. In the eastern and central equatorial Pacific upwelled waters are cold

14 TAO/TRITON Monthly SST ('C) and Winds (m s-l) (a) 1400E 160"E 180' 160'W 120'17 100' W lOON 31 2. S'N 27 0' 25 23 5'S 21 10'S 1I.

December 1997 An omalies

tolay JO 2007

TAO/TRITON MonthLy SST ('C) and Winds (m .-1)

(b) 1400E 160'E 120'17 100'W 10'N .,-~~-~ 31 S' N 2. 27 D· 25 23 5'S 21 10'S '-'='JIoo,--+---'--~-=F~ 119 ~ 10. m e-1

1~ , 1;~~ C!l~=-::==~~~~~~g;~~~~~~C;:~;;~1 1 ~ December 1998 Anomalies

TAO/TRITON Mon th Ly SST ('C) and Winds (m .-' )

(c) 140'E 100'W lOON 31 2. S' N 27 0' 25 23 5'S 2' 10 0 S ,. May 2005 Means

TN:! ProJ~t OIrlc./ PIotEL/NCWI tolay ZIt 2007

Figure 6. Mean monthly sea surface temperatures, winds and anomalies between 10' S and 10' N from the Tropical Atmosphere Ocean (TAO) arrays located in the equatorial Pacific. (a) EI ino, December 1997, (b) La ina, Dec 1998 and (c) Normal, May 2005.

15 and rich in nutrients leading to HNLC conditions while in the west the upwelled waters are nutrient-poor or oligotrophic. The transition from the warm pool to the HNLC waters is characterized by a marked increase in surface salinity (picaut et aI., 1996), an increase in the concentration of surface nutrients, a decrease in temperature and an increase in the surface partial pressure of C02 (Inoue et a!., 1996; Stoens et aI., 1999; Rodier et al.,

2000; Le Borgne et al., 2002). The location of the front between the warm pool and

HNLC water varies greatly depending on whether EI Nifio or La Nifia climatic conditions prevail (Inoue et al., 1996; Picaut et al., 1996; Delcroix and Picaut, 1998). This study transected the equator at approximately 167" W in April-May 2005, in the HNLC waters of the central equatorial Pacific under normal climatic conditions in which cooler sea surface temperatures occur in the eastern equatorial Pacific due to upwelling and the anomaly showed very little change from mean c1imatology (Figure 6c).

2.1.4 Slze-/ractionation

In general, methods for measuring total particulate material in aquatic ecosystems use filters while the methods for measuring size fractions of the particulate material use screens (Sheldon, 1972). Sheldon (1972) compared particle removal by cellulose

(MiIlipore) and glass-fiber (Wbatman) filters with particle removal by polycarbonate membrane (Nuclepore) filters. He found that polycarbonate membrane filters produce the sharpest separation of particles into size classes smaller than their pore size. Glass­ fiber and cellulose filters removed particles over a much wider size interval and were not recommended for separating particles by size. In the literature there has been extensive debate on Chi a retention properties of glass-fiber fine (GFIF; Wbatman) filters and

16 0.2 J.LIIl membrane filters (Dickson and Wheeler, 1993; Chavez et al., 1995). Dickson and

Wheeler (1993) reported significantly bigher concentrations ofCbl a for 0.2 J.LIIl membrane filters compared to GFIF filters. However, based on several independent open ocean comparisons of these two filter types, Chavez et al. (1995) found contradictory results to Dickson and Wheeler (1993), showing little or no difference between their retention properties. Throughout this study, total particulate material was collected using

GFIF (Whatman, nominal pore size 0.7 J.LIIl) filters and size-fractionated particulate material was filtered using polycarbonate membrane (Nuclepore) filters with pore diameters of20, 10, 2 and 0.2 J.LIIl as a way of separating phytoplankton assemblages into various size groups.

2.1.5 Objectives ofthis study

The size structure and taxonomic composition of the phytoplankton community in the open ocean are important factors in regulating organic carbon export to the deep ocean

(Azam et al., 1983; Tremblay and Legendre, 1994). The main objective of this study was to understand the spatial variability of the phytoplankton size structure and community composition across biogeochemical provinces in terms of nutrient availability. Here I present the spatial distributions of surface nutrients, phytoplankton size structure, abundance and taxonomic composition from one cruise within the Pacific Ocean. The cruise transected nutrient gradients within and across the south Pacific, central equatorial

Pacific and north Pacific waters and encompassed four biogeochemical provinces as defined by Longhurst (1998).

17 2.2 METHODS

2.2.1 BEACH-BASH Data

Sampling was conducted from the RN Kilo Moana during, BEACH-BASH, a latitudinal transect (12·S - 22·N) cruise across the Pacific Ocean from Pago Pago, American Samoa to Honolulu, Hawaii in May of2005. A total of28 stations were occupied during the transect (Figure 3). The stations were spaced at approximately equidistant intervals, 1_2° apart starting at 12°41.4'S latitude, 1700 26.5'W longitude. The last station occupied coincided with the Hawaii Ocean Time-series (HOT) Station ALOHA (22°45'N, 158°W)

(Karl and Lukas, 1996). Hydrographic data (temperature, salinity, pressure, density, fluorescence and oxygen) and in situ water samples were acquired from the near surface

(-5 m depth) ocean using either a SeaBird SBE 9/11 + CTD system plus rosette sampler fitted with 24 x 12 L Niskin-type water bottles or from the hull-mounted underway uncontaminated seawater intake system positioned at 7 m. The CTD fluorometer was calibrated against Chi a concentrations extracted from water samples and determined by fluorometric analysis. Current velocities were measured between the 25 m to 35 m depth layer using an RD Instruments 300 kHz WorkHorse Acoustic Doppler Current Profiler

(ADCP), mounted in the port hull of the ship at approximately 7 m.

2.2.2 Sample coUection and analysis

2.2.2.1 Nutrientanalysis

Water samples for nutrient analysis (nitrate + nitrite [N+N], soluble reactive phosphorus

2 [SRP] and silicic acid [Si(OH)4- ]) were collected in acid-washed 125 ml or 500 ml high density polyethylene (HOPE) bottles, immediately frozen (-20· C), and stored upright

18 until analyzed. Because nitrogen and phosphorus dynamics in the oligotrophic surface waters of the North Pacific Ocean are known to have ambient concentrations typically at or below the detection limits of standard colorimetric methods (Karl, 1999 and 2002), high-sensitivity analytical methods were used to determine surface concentrations of

N+Nand SRP.

Surface N+N determinations were perfonned using an Antek model 720 chemiluminescence nitrogen analyzer following the technique of Cox (1980) as modified for seawater by Garside (1982). This technique relies on the wet chemical reduction of nitrate (N03) and nitrite (NCh) to nitric oxide (NO) using ferrous ammonium sulfate, ammonium molybdate and a highly acidic solution of sulfuric acid. The limit of detection for N+N was approximately 0.001 pM with a triplicate precision and accuracy of ± 3 % (Table 1).

Surface SRP concentrations were analyzed using the magnesium induced co-precipitation

(MAGIC) procedure (Karl and Tien, 1992), which has been used extensively for analysis at the HOT site (Thomson-Bulldis and Karl, 1998; Karl and TieD, 1997). Frozen samples were thawed in a water bath and brought to room temperature. Subsamples for MAGIC

(3 replicates, 50 ml each) were decanted into disposable 50 ml polypropylene centrifuge tubes. A high purity NaOH solution was added at 1 % (NaOH/sample; vol/vol basis) to fonn the mineral brucite Mg(0H)2. Samples were than capped, inverted to mix the flocculate produced by NaOH and centrifuged at 1000 x g for 60 minutes. During this process the insoluble Mg(OHn containing the adsorbed phosphorus forms a pellet.

19 Table 1. Detection limit and precision estimates, volume sampled and observed ranges for all assays during the BEACH-BASH transect.

Detection limits & AssaI Precision Volume Values observed P-ATP GFIF 5ngll± 12% IL 15 - 87 ngll >20 !UD 5ngll± 12% IL 20 !UD O.Olugll ± 20010 IL 0.026 - 0.038 ugIl 10-20!UD 0.0 I ugIl ± 20% I L 0.027 - 0.031 ugIl 2-10!UD 0.0 I ugIl ± 20% IL 0.027 - 0.038 ugIl 0.2 - 2!UD 0.0 I ugIl ± 20% 155 ml 0.052 - 0.215 ugIl Phycoerythrin >20 !UD 0.5ngll ± 30010 5L < DL - 0.81 ngll 10-20 !UD 0.5ngll ± 30010 5L < DL - 1.16 ngIl 2-10!UD O.5ngll ± 30010 5L < DL - 1.73 ngll 0.2-2!UD 0.5ngll ± 30010 IL 0.70-31 ngll Floweytometry 4 4 Heterotrophic bacteria N/A± 15% 2m! 52 x 10 - 88 x 10 cells mr' 4 4 Procblorococcus spp. N/A± 15% 2m! 1.5 x 10 - 15 X 10 cells m!.' 4 4 Synechococcus spp. N/A± 15% 2ml 0.04 x 10 - 1.88 X 10 cells mr' 4 4 Picoeukaryotes N/A± 15% 2m! 0.08 x 10 - 0.78 X 10 cells mr' HPLC Total cbloropbyU a* NI A ± 2.0010 4L 58.67 - 275.98 ngll Monovinyl cbloropbyU a N/A±0.83% 4L 29.32 - 146.96 ngIl Divinyl cbloropbyU a N/A±0.75% 4L 21.08 - 119.60 ngIl Monovinyl cbloropbyUide a NlA±0.17% o- 34.90 ngll Monovinyl plus divinyl cblorophyll b N/A±0.33% 4L 1.51 -26.68 ngll Chlorophyll c-like pigments N/A± 1.9% 4L 6.59 - 55.27 ngll 19' - Butanoyloxyfucoxanthin N/A±0.22% 4L 3.00 -35.38 ngll 19' - Hexanoyloxyfucoxanthin NI A ± 0.30010 4L 9.89 - 69.94 ngIl Fucoxanthin N/A±0.13% 4L 1.75 - 22.09 ngIl Peridioin N/A±0.93% 4L 0- 8.82 ngll Prasinoxanthin N/A±029% 4L 0 VioIaxanthin N/A±0.68% 4L 0- 2.20 ngll Diadinoxanthin N/A± 1.61% 4L 3.01 - 34.01 ngIl Lutein N/A±0.23% 4L 0- 3.66 ngll Alloxanthin N/A±O.46% 4L 0 Zeaxanthin N/A±0.71% 4L 44.34 - 87.58 ngll p,& - Carotene N/A±0.77% 4L 4.21 - 18.28 ngIl p,p - Carotene N/A±0.76% 4L 1.26 - 6.68 ngll Nutrients Nitrate + Nitrite (N+N] O.00IpM±3% 50m! 0.0025 - 3.7 pM Silicic acid [Si(OH).-2] 0.lpM±5% 50m! 0.55 - 2.65 pM Soluble Reactive PhOSJ)horus [SRP] 0.003pM±2% 50m! 0.06 - 0.43 pM

20 Following centrifugation, the supernatant was aspirated using a Pasteur pipette attached by Tygon@tubingto a vacuum line, being careful not to disturb the small pellet. The pellet was then dissolved in 5 ml of high purity O.IM HCI. The arsenate reduction mix was then added to each 5 ml sample in aliquots of 0.5 ml and allowed to react for 15 minutes. After 15 minutes 0.5 ml of the molybdenum blue reaction mix was added to each 5 ml sample, mixed and allowed to react for another 15 minutes to ensure full color development Samples were then read at 880 nm in a 10 em cell using a Beckman DU

640 Spectrophotometer. Typical precision estimates for triplicate determinations of SRP were ± 2 % with a detection limit of 0.003 J.LM (Table 1).

Triplicate surface Si(OH)4·2 concentrations were measured spectrophotometrically. This analysis was based upon the formation of yellow silicomolybdic acid from the reaction of ammonium molybdate and silica at acidic pH. As modified from Strickland and Parsons

(1972), the sensitivity of the analysis was increased by a further reduction of the yellow silicomolybdic acid using ascorbic acid. in order to produce "molybdenum blue." The silicomolybdenum blue complex was then measured at 810 nm in a 10 em cell using a

Beckman DU 640 Spectrophotometer. The limit of detection for Si(OH)4-2 was approximately 0.1 J.LM with a triplicate precision and accuracy of ± 5 % (Table 1).

2.2.2.2 Chlorophyll a

Surface Chi a concentrations were determined by both standard fluorometric techniques

(Holm-Hansen et al.• 1965; Strickland and Parsons, 1972) and by high-pressure liquid

chromatography (HPLC; Wright et al., 1991; Bidigare et al., 2005). The later method is

21 described in further detail in section 2.2.2.6. Samples for tluorometric analysis of total

Chi a (TChI a) were made by filtering 155 ml of seawater through a 25 mm diameter glass fiber filter (Whatman GFIF, nominal pore size 0.71=) via vacuum filtration.

Surface size-fractionated Chi a samples were measured by sequentially filtering 1 liter of seawater through a 25 mm diameter polycarbonate membrane filter (Nuclepore) with pore diameters of20, 10,2 and 0.21= via pressure filtration. Immediately following filtration, the filters were transferred to a glass screw cap tube containing 5 ml of cold

(-20'C) 100 % acetone. The tubes were then wrapped in aluminum foil and stored at

-20'C to prevent photodegradation of pigments. Chi a concentrations were measured by tluorometric analysis on a Turner Designs TD-700 with 436 nm (Ex) and 680 nm (Em) wavelength filters.

2.2.2.3 Phycoerythrin

Samples for the measurement of and photosynthetic accessory pigments are extracted in acetone (see section 2.2.2.6) and therefore select against phycoerythrin, a water soluble pigment which can be used as a biomarker for cyanobacteria such as

Synechococcus spp. and Trichodesmium spp. Surface seawater samples for phycoerythrin analysis were collected in a 5 L carboy and size fractioned sequentially through a 25 mm diameter polycarbonate membrane filter with pore diameters of20, 10, and 2 1= via pressure filtration following the in vivo glycerol-uncoupling method of

Wynam (1992) as modified by Dore et al. (2002). One liter of filtrate was then collected and passed through a 25 mm diameter 0.2 1= polycarbonate membrane filter. Filters were placed in 20 ml glass scintillation vials containing 5 ml of a saline 50 % glycerol

22 solution. The filters were placed up against the inner wall of the vial so that particles from the filters could become suspended in the solution during gentle shaking.

Phycoerythrin concentrations were measured by fluorometric analysis using a Turner

Designs TD-700 with 544 nm (Ex) and 577 nm (Em) wavelength filters. The TD-700 fluorometer was calibrated using buffered solution of a commercial R-phycoerythrin standard (Cyanotech Corp.).

2.2.2.4 Adenosine 5 '-triphosphate (ATP)

The amount of living microbial biomass in the surface water was determined by the measurement of particulate adenosine 5' -triphosphate (P-A TP) concentrations (Holm­

Hansen and Booth, 1966). Seawater samples for total and size-fractionated P-A TP were passed through a 202 IJ.IIl mesh Nitex® screen to remove large plankton and collected in

lUter carboys. Surface total P-A TP samples were filtered onto 47 mm diameter GFIF filters. Surface size-fractionated P-ATP samples were collected and filtered through a

47 mm diameterpolycarbonate membrane filter with pore diameters of20, 10,2 and

0.2 IJ.IIl via vacuum filtration. The filtrate from each size fraction was then collected and filtered through 47 mm diameter GFIF filters. Upon completion of filtration P-ATP was extracted by placing the GFIF filters in 5 m1 of boiling TRlS buffer (20 mM, pH 7.4) for

S. minutes. Extracted ATP concentrations were then analyzed in a Turner Biosystems

20120· Luminometer using the firefly bioluminescence reaction (Karl and Holm-Hansen,

1978; Karl, 1993). To estimate the contribution of each size fraction to the total P-A TP pool, the P-ATP concentrations in individual fractions were calculated by difference.

Because of unidentifiable errors in size fractioning and calculating P-A TP concentrations

23 in each fraction by difference the absolute P-ATP concentrations in these operationally defined size classes should be viewed as estimates. The relative changes and spatial trends in P-ATP concentrations for each size class provide infonnation on the distributions of microbial assemblages across the BEACH-BASH transect.

2.2.2.5 Flow Cytometry

Surface flow cytometry samples were collected for enumeration ofpicoplankton

(Prochlorococcus spp., Synechococcus spp., and picoeukaryotes) and heterotrophic bacteria. Seawater samples were taken in 2 ml aliquots, placed into cryoviaIs and preserved using 40 !Jl of 0.5 )Ill1 filtered 10 % paraformaldehyde. Cryovials were then stored in a -80°C freezer until analyzed on the flow cytometer. Samples were thawed in batches, then stained with Hoechst 33342 (1 JLglmJ, final concentration) (Monger and

Landry, 1993; Campbell and Vaulot, 1993; Campbell et al., 1994). Anintemal standard to normalize scatter and fluorescence signals was added to each sample. The standard was a mixture of anto-fluorescent polystyrene beads, consisting of 0.5 JLm diameter yellow-green fluorescence beads, 0.5 JLm diameter UV beads and 1 JLm yellow-green fluorescence beads. The flow cytometer used was a Beckman-Coulter A1tra, mated to a syringe pump for quantitative analyses, and was equipped with two argon ion lasers, tuned to UV (225 mW) and 488 run (1 W) excitation. Scatter (side and forward) and fluorescence signals were collected using filters as appropriate, including those for

Hoechst-bound DNA, phycoerythrin and Chi a. Population designations, based on the scatter and fluorescence signals, were generated from the Iistmode files using FlowJo software (Tree Star, Inc., www.flo~o.com).

24 2.2.2.6 HPLC Pigments

Surface water samples for the measurement of chlorophylls and taxonomic biomarker accessory pigments were collected from 24 stations during the BEACH-BASH transect starting from 6.440 S. Samples were collected in 4 liter HOPE bottles and pressure filtered through 25 rom GFIF filters. The filters were placed in cryovials, flushed with N2 gas, and stored in liquid nitrogen until analysis. Pigments were extracted and analyzed by reverse-phase high performance liquid chromatography (HPLC; Wright et a/., 1991;

Bidigare et a/., 2005). Filters for pigment analysis were extracted in 3 ml of HPLC-grade acetone containing a known amount of an internal standard (cantbaxanthin) at 4"C for 24 hours. For each sample mixtures of 1 ml extract (containing cantbaxanthin) and 300 J1l

HPLC grade water were injected into the autosampler of the Varian 9012 HPLC system.

Photosynthetic pigments were separated on a reverse-phase Waters Spherisorb@ 5 !1111

ODS-2 (4.6 x 250 rom) CIS column with a corresponding guard cartridge (7.5 x 4.6 rom) and a Timberline column heater (26"C). Pigment identifications were based on absorbance spectra, co-chromatography with standards and relative retention time.

Abbreviations and significance of photosynthetic pigments discussed in this study are described in Table 2.

To facilitate a broad assessment of the relative importance of two functional classes of phytoplankton along the BEACH-BASH transect a diagnostic pigment (DP) index

(Bidigare pers corom.) was defined as the total of eight pigments selected for their taxonomic affinity: [DP ()LgIl) = Zea + DVChl a + TChl b + a-Car + Hex + But + Fuco +

TChl c], where zeaxanthin (Zea) is indicative of cyanobacteria, divinyl chlorophyll a

25 Table 2. Abbreviations and taxonomic affinities ofphotosyntbetic pigments separated in this study using high performance liquid chromatography (HPLC).

Pigment name Abbreviation Primary pigment in:

Total chlorophyll a* TChla All phytoplankton Monovinyl chlorophyll a MVChla All phytoplankton (except Prochlorococcus spp.) Divinyl chlorophyll a DVChla Prochlorococcus spp. Monovinyl chlorophyllide a MVChlda Senescent diatoms Monovinyl plus divinyl chlorophyll b TChlb Prochlorococcus spp. Chlorophyll c-like pigments TChlc Chromophytes 19' - Butanoyloxyfucoxanthin But Pelagophytes 19' - Hexanoyloxyfucoxanthin Hex Prymnesiophytes Fucoxanthin Fuco Diatoms Peridinin Per Dinoflagellates Prasinoxanthin Pras Prasinophytes Violaxanthin Viola Chrysophytes Diadinoxanthin DDX Chromophytes Lutein Lut Chlorophytes Alloxanthin Allox Cryptophytes Zeaxanthin Zea Cyanobacteria 13,e - Carotene a-Car Prochlorococcus spp., cryptophytes 13,13 - Carotene j3-Car All phytoplankton groups (except Prochlorococcus spp.)

>I< TChl a = MVChl a + DVChl a + MVChld a Pigment description form Jeffery et aI. (1997)

26 (DVChl a), total chlorophyll b (TChl b), and alpha carotene (a-Car) of Prochlorococcus spp. and prochlorophytes, 19' -hexanoyloxyfucoxanthin (Hex) of prymnesiophytes,

19'-butanoyloxyfucoxanthin (But) of pelagophytes, fucoxanthin (Fuco) of diatoms and total chlorophyll c (TChl c) of chromophytes. The proportion ofDP attributed to each of the functional classes of phytoplankton is defined as:

• [Zea + DVChl a + TChl b + a-Car] / DP x 100 or [CyanoDP] -used to indicate

the proportion of prokaryotic picoplankton i.e. cyanobacteria and

prochlorophytes.

• [Hex + But + Fuco + TChl c) / DP x 100 or [ChromoDP] - used to indicate the

proportion of chromophyte microalgae i.e. prymnesiophytes, pelagophytes, and

diatoms.

While these identified groups must be viewed with caution they have been chosen based on the community structure found in previous studies, in which cyanobacteria and prochlorophytes are known to contribute to phytoplankton biomass in oligotrophic waters

(Chisholm, 1992; Campbell and Vaulot, 1993) and chromophyte microalgae such as prymnesiophytes, pelagophytes and diatoms are known to contribute to phytoplankton biomass in the Pacific equatorial waters (Bidigare and Ondrusek, 1996). The grouping of different diagnostic pigments has been used in other studies to look at the size and taxonomic composition of the phytoplankton community (Vidussi et al., 2001; Barlowet al., 2004; Poulton et aZ., 2006).

27 2.3 RESULTS

2.3.1 Habilllt Characteristics

2.3.1.1 Biogeochemical provinces

ADCP current velocities between the 25 m to 35 m depth layer are shown in Figure 5, and representative vertical distributions of temperature, fluorometric TChi a and SRP from each province are shown in Figure 7. During the BEACH-BASH cruise the SPSG province was sampled starting from 12°41.4'S and encompassed a total of 5 stations

(Figures 3 and 4). The vertical distribution of temperature, TChi a and SRP in the SPSG province are shown in Figure 7a The SPSG exhibited a mixed layer depth of approximately 75 m based on a 0.125 unit change in potential density. Surface temperatures were approximately 30' C and decreased with increasing depth to a temperature of22' C at 200 m. TChi a concentrations were 0.10 IJ.gI1 at the surface and exhibited a subsurface maximum of 0.30 IJ.gI1 at the base of the mixed layer depth (75 m).

Surface water SRP concentrations were fairly low (0.14 )JM) but increased below the mixed layer to a maximum concentration of 0.66 !1M at a depth of 150 m.

The surface circulation in the PEQD province was dominated by the flow of the SEC which moved westward across the ocean and the region of divergence located at the equator (Figure 5). During the BEACH-BASH cruise the PEQD province encompassed a total of 12 stations (Figures 3 and 4). The vertical distribution of temperature, TChi a and SRP in the PEQD province are shown in Figure 7b. The PEQD exhibited a mixed layer depth of approximately 55 m. Surface temperatures were 28" C and decreased with increasing depth reaching a minimum temperature of 14° C at 200 m. TChi a 28 SPSG PEQO 0 0 (a) (b)

50 Ul 50 ~ nl .0 ~ ~ 100 100

!/)=> !/) Q) ~ a.. 150 150

200 200

PNEC NPTG 0 0 (c) (d --- 50 50 Ul ~ nl .0 ~ ~ 100 100 => !/) !/) Q) ~ a.. 150 150

12 14 16 18 20 22 24 26 28 30 12 14 16 18 20 22 24 26 28 30 Temperature (0G) Temperature (OC)

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 Chlorophyll a (I'!I/I) Chlorophyll a h'9/1) ,- - ,- .---r---,---~--~--~' ~ 0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2

SRP (~M) SRP (~M)

Figure 7. Depth profiles of temperature (blue), fluorometric Chi a (green), and soluble reactive phosphorus (SRP) (red) at representative stations in the biogeochemical provinces sampled during the BEACH-BASH transect cruise. The dashed line indicates the mixed layer depth based on a 0.125 unit change in potential density. (a) SPSG (10.51 ° S), (b) PEQD (0.63° N), (c) PNEC (6.00° N) and, (d) NPTG (J 5.85° N).

29 concentrations at the surface were 0.15 p.g/l and exhibited a subsurface TChl a maximum of 0.30 p.g/l at 60 m. Surface SRP concentrations were 0.45 JlM. These concentrations were higher than those found in the SPSG province and steadily increased with depth to a maximum concentration of 1.0 JlM at 175 m.

In this PNEC province the curvature of wind stress not only forces the flow of the eastward NECC, but is also thought to enhance algal growth and therefore Chl a concentrations (Longhurst, 1993; Figmes 3 and 5). During the BEACH-BASH cruise the

PNEC province encompassed a total of 4 stations (Figmes 3 and 4). The vertical distribution of temperature, TChl a and SRP in the PNEC province are shown in

Figure 7c. The PNEC exhibits a mixed layer depth of approximately 100 m. Surface temperatures of 29" C extended throughout the mixed layer (100 m) with temperatures reaching a minimum of 12° C at a depth of200 In. TChl a surface concentrations were

0.28 J,tgll. These concentrations were higher than those found in the PEQD province and extended down to a depth of 100 m. No subsurface TChl a maximum occurred. Surface

SRP concentrations were 0.2 JlM. These concentrations were lower than those found in the PEQD province. SRP concentrations increased below the mixed layer depth to a maximum concentration of 1.2 JlM which occurred at a depth of 175 m.

The circulation in the NPTG province between 10' N and 15' N was dominated by the westward flowing North Equatorial Current (NEC; Figure 5). During the BEACH-BASH cruise the NPTG province encompassed a total of 6 stations with the last station coinciding with Station ALOHA at 22°45' N (Figmes 3 and 4). The vertical distribution

30 of temperature, TChl a and SRP in the NPTG province are shown in Figure 7d. The

NPTG exhibits a shallow mixed layer depth at approximately 35 m. Surface temperatures were approximately 26° C and decreased steadily with increasing depth to a temperature of ISO C at 200 m. TChl a concentrations were 0.07 !LW'l at the surface and exhibited a deep subsurface maximum of 0.27 !LW'l at a depth of 150 m. SRP concentrations in surface waters were fairly low (0.15 !J.M) and remained low throughout the upper 175 m.

2.3.1.2 Latitudinal distribution o/Temperature, Salinity and Chlorophyll a

The latitudinal distribution of surface temperature, salinity and fluorometric TChl a are shown in Figures Sa and b. In general, surface temperatures decreased from 30· C at the southern extent of the transect to 25.5" C at the northern end of the transect. At the equator there was an abrupt drop in temperature to 2S· C followed by a rapid increase in temperature just northward (Figures Sa and b). Surface salinity ranged from 34.25 to

35.5 Practical Salinity Scale (PSS; Figure Sb). Surface salinity was highest at the southern and northern most ends of the transect as well and just south of the equator. The sea surface temperature was relatively low and the salinity relatively high in the equatorial area which indicated upwelling. Fluorometric TChl a values ranged from 0.06 to 0.28 !LW'l with the highest concentrations occurring in the equatorial provinces (PEQD and PNEC; Figure Sb). TChl a concentrations peaked north of the equator in the PNEC province and were followed by an abrupt drop in concentration from 0.2S to 0.08 !LW'l.

TChl a concentrations in the SPSG province (0.07 to 0.10 !LW'l) were slightly higher than

TChl a concentrations measured in the NPTG province (0.06 to 0.09 !LW'l).

31 Surface Temperature, Salinity and Chlorophll a SPSG PEQD PNEC NPTG 31 -.------="----='-=--..,-----'---= -=-----.,.-'--'c..:=.=-r---'-.:..:..--'-=------, 0.30 (a) -- Temperature 30 - Chlorophyll a 0.25

~ ~ ()o 29 " ~ ~ ~ 0.20 -; :::l >­ "§ 28 .s:::. Q) a. a. 0.15 E eo ~ 27 .s:::. () 0.10 26

25 +-----.--I-----,---+--t---,-----.--+ 0.05 31 36.0 (b) -- Temperature 30 -- Salinity 35.5

oG' 29 ~ if) (/) ~ a.. :::l ~ 'iii 28 35.0 ~ ~ Q) a. c: ro E (/) ~ 27 34.5 26

25 +-- ---.--I-----,---+--t---,-----.--+ 34.0 -15 -10 -5 o 5 10 15 20 25 Latitude (Oeg)

Figure 8. Latitudinal distribution of surface (a) temperature (0C) and chlorophyll a (!!g/l) and (b) temperature (OC) and salinity (PSS).

32 2.3.2 Nutrient distributions

2.3.2.1 LatitudiTUll distribution ofdissolved iTUlrganic nutrients

The latitudinal distribution of surface inorganic nutrients N+N, SRP and Si(OH)4"2 are shown with their standard deviations in Figure 9. In most cases standard deviations were less than or equal to the size of the data point and are not visible. All nutrient concentrations exhibit a pronounced increase at the start of the PEQD province (5° S) which coincided with the start of the strong westerly SEC and region of equatorial divergence (0") where nutrients are upwelled to the surface layer from below the thermocline (about 75 m, refer to Figure 7b). N+N concentrations ranged from

0.0025 - 3.73 JIM throughout the transect. Consistently low N+N values were found in

SPSG, PNEC and NPSG provinces. N+N concentrations increase sharply at the start of the PEQD province (5° S) and remain relatively high until reaching the equator where a pronounced drop in N+N concentration occurred. N+N concentrations exceeded SRP concentrations between 5° S and 0°. Surface SRP concentrations ranged from

0.06 - 0.43 JIM throughout the transect. SRP concentrations steadily increased at the start of the PEQD, peaked at the equator and then steadily decreased through the PNEC.

SRP concentrations were lowest in the SPSG and NPTG provinces but exceeded N+N concentrations in the SPSG, PNEC and NPTG province. Surface Si(OH)4"2 concentrations ranged between 0.55 - 2.65 JIM throughout the transect and followed the same pattern as N+N. Surface Si(OH)4"2 concentrations were lowest in the SPSG and increased sharply at the start of the PEQD province, remaining high until reaching the equator where a pronounce drop occurred. Si(OH)4"2 concentrations in the PNEC and

NPTG provinces were greater than those found in the SPSG province.

33 Nutrients SPSG PEQO PNEC NPTG 4 2.0 - N+N - SRP ~ 3 1.5 :2 - Silicic Acid :::1. ~ -0 Tj ~ rn 2 1.0 :2 u :::1. ~ : ~ a.. en ...... ~ 0::: o(l 1 0.5 en z + ~- z 0 0.0

-15 -1 0 -5 o 5 10 15 20 25 Latitude (Oeg)

Figure 9. Latitudinal distribution of surface nutrients nitrate + nitrite [N+N], soluble 2 reactive phosphorus [SRP] and silicic acid [Si(OHk ] along the BEACH-BASH transect.

34 2.3.2.2 Dissolved inorganic nutrient ratios

The dissolved inorganic N+N : SRP ratios in the surface waters throughout the transect never exceeded the Redfield ratio 16:1 (Redfield et aI., 1963; Figure lOa). The

N+N : SRP ratios were highest in the PEQD province between 5° S - 0° (mean = 7.6

± 1.18, n = 6), where the highest N+N concentrations were observed. The N+N : SRP ratios were significantly lower in the SPSG, PNEC and NPTG provinces (mean = 0.08, n = 16).

The dissolved inorganic Si(OH),,2 : N+N ratio in the surface waters throughout the transect almost always exceeded the 1: 1 ratio of growing diatoms (Brzezinski, 1985), except in the PEQD province between 5° S - 0°, where equatorial upwelled waters were observed (Figure lOb). In this region the ratio of Si(OH)4·2 : N+N approached 1: 1

(mean = 0.68 ± 0.09, n = 6). The Si(OH)4·2 : N+N ratios dramatically increased between the equator and 5° N exhibiting a maximum of 500 at approximately 4° N. The mean

Si(OH)4·2 : N+N ratios in the SPSG, PNEC and NPTG provinces were 90, 135 and 185 respectively.

The dissolved inorganic Si(OH)4·2 : SRP ratio in the surface waters throughout the transect almost never exceeded the ratio 16: 1 of growth, except in the northern most Station ALOHA (Si(OH)4·2 : SRP = 18; Figure lOe). The Si(OH)4·2 : SRP ratio gradually increased along the transect from south to north. The mean Si(O~'2 : SRP ratio in the SPSG, PEQD, PNEC and NPTG provinces were 4, 4.5, 7 and 12.5 respectively.

35 SPSG PEQD PNEC NPTG 100 (a) ------f------10 ...... a.. a:: en 1 • • Z • • + 0.1 1 • • •• ~ • • Z • •• • • • • • 0.01 •• 0.001 -15 -10 -5 o 5 10 15 20 25

1000 (b) •••• • Z •• ~ • • + 100 , • . • Z • • • • • "0 '(3 10 ttl • • U : ~ • • en 1 c------.-:••••------

0.1 -15 -10 -5 o 5 10 15 20 25

100 (c) a.. a:: en f------"0 '(3 10 , • • • • • ttl • • U • •• :~ • ., ... ' .... Ui • •• • •

1 -15 -10 -5 o 5 10 15 20 25 Latitude (Oeg)

Figure 10. Latitudinal distribution of surface nutrient ratios; (a) ratio ofN+ to SRP, with dashed line at a ratio of 16: I, (b) ratio of silicic acid to N+N, with dashed line at a ratio of I: I, (c) ratio of si licic acid to SRP, with dashed line at a ratio of 16: I across BEACH-BASH transect. 36 2.3.3 Plankton Size Structure

2.3.3.1 Chlorophyll a

The latitudinal distributions of fluorometric surface size-fractionated ChI a concentrations are shown in Figures lla and b. Highest levels of phytoplankton ChI a were observed in waters located in the PEQD and PNEC provinces between 5' S and

10' N with a peak in all size fractions occurring in the PNEC province at approximately

6' N. The latitudinal distribution of ChI a was dominated by picoplankton (0.2-2 lUll) in all provinces (Figure II a). Picoplankton Chi a closely reflected the patterns of abundance of total phytoplankton, given that this fraction accounted for 45 - 60 % of the

TChi a during most of the transect (Figure 11 c). Lowest levels of phytoplankton ChI a were observed in waters located in the oligotrophic SPSG and NPTG provinces with picoplankton values below 0.1 )1gI1. Flow cytometry analysis indicated that the cyanobacteria Prochlorococcus spp. and Synechococcus spp. constituted the bulk of the picoplankton abundance during the transect (Figure 17).

Nanoplankton (2-10 lUll and 10-20 lUll) Chi a concentrations were markedly lower than those of picoplankton. Typical surface values ranged between 0.027 - 0.036 )1gI1, with nanoplankton of 2-1 0 lUll in size having slightly higher concentrations than nanoplankton of 10-20 lUll in size throughout the transect. This can be seen most noticeably in the

PEQD and PNEC provinces (Figure lib). The highest concentration of nanoplankton

ChI a (2-10 lUll) was 0.038 )1gI1 and occurred in the PNEC province at 6' N. The highest concentration of nanoplankton ChI a (10-20 lUll) was 0.031 )1gI1 and occurred in both the

PEQD and PNEC provinces. Each of the nanoplankton size fractions typically

37 Size - Fractionated Chlorophyll a 0.25 -,----=-==----SPSG -,----'PEQD'-==--r'- PNEC:....==-, ----'--'''-'-=-NPTG ----, (a) -- > 20 I'm ::::::~ 0.20 -- 10-20 I'm ~ -- 2-10 I'm ~ (\) 0.15 -- 0.2 -21'm >- .r:: a. 0 0.10 ~ 0 .r:: • .... ---- - () 0.05 . •

0.00 0.040 (b) -- >20 I'm ~ -- 10- 20 l'm ""~ 0.036 -- 2 -10 I'm ~ (\) 1= 0.032 a. 0 ~ 0 .r:: 0.028 ()

0.024 -15 -10 -5 o 5 10 15 20 25 Latitude (Deg) 70 (C) _ >20 I'm 60 _ 10- 20 l'm (\) _ 2- 10 l'm >. .r:: 50 _ 0.2 -2 I'm a. 0 ~ 0 40 .r:: () 30 til 0 I-- 20

0~ 10

0 SPSG PEQD PNEC NPTG Province Figure II . Latitudinal di stribution of surface size-fracti onated chlorophyll a (a) >20 I!m, I 0-20 ~ , 2- \0 I!m and 0.2-2 I!m (b) >20 I!m, 10-20 I!m, and 2-1 0 I!m, and (c) Regionalized percent (%) total surface chloropbyll a.

38 represented 15 - 20 % of the TCbl a throughout the transect (Figure llc), therefore total nanoplankton (2-20 11IIl) represented 30 - 40 % of the TCbl a.

Microplankton (>20 11IIl) Cbl a concentrations were also markedly lower than those of picoplankton and in most cases lower than both nanoplankton (2-10 and 10-20 11IIl) size fractions (Figure 11 b). Microplankton Cbl a was present in surface waters with typical concentrations ranging between 0.026 - 0.030 j.tgIl. Concentrations of microplankton

Cbl a peaked in the PEQD province at approximately 4° N with a value of 0.038 p.g/l.

Microplankton Cbl a typically represented 12 - 18 % of the TChI a throughout the transect (Figure Ilc). TCbl a measurements made using glass fiber filters (GFIF) were well correlated with TCbl a measurements made using the sum of the polycarbonate filters (PC); (pC = 0.82 (GFIF) + 0.083, ~ = 0.987, p < 0.001, n = 27; Figure 12).

Correlations between ChI a size fractions and nutrients were observed (Figures 13a, b and c). N+N showed the weakest correlation to all Cbl a size fractions with Pearson correlation coefficients ranging berween r = 0.077 - 0.334 (Figure 13a). Microplankton

Cbl a (>20 11IIl) appeared only weakly dependent on N+N (r = 0.077) while nanoplankton

Cbl a (10-20 11IIl) demonstrated greater dependence (r = 0.334). Similar to the ChI a relationships to N+N, nanoplankton Cbl a (10-20 11IIl) concentrations appeared most strongly dependent on Si(OH)4·2 and SRP concentrations (r = 0.594 and r = 0.732, respectively; Figures 13b and c).

39 0.35 -,------, • 0.30

E" ~ 0.25 • • :c ~u 0.20 Q. •

0.15 y =0.82 x + 0.083, R2 =0.987 0.10 +-----,---,----.------r----j 0.05 0.10 0.15 0.20 0.25 0.30 GFIF TChl a (Ilg/l)

Figure 12. Chlorophyll a regression analysis. TD-700 fluorometric TChl a measured using GFIF filters versus TChl a measured using summed PC filters.

40 •

(a)

~ Ol '"::t ~ Total Chi a (r = 0.294) (1) 0.1 • > 20 ~m Chi a (r = 0.077) >- .c • '" 10-20 ~ m Chi a (r = 0.334) a. 0 .II. 2-10 ~ mChla(r=0 . 198) ~ 0 :c • 0.2-2 ~m Chi a (r = 0.179) u

0.01 0.001 0.01 0.1 1 10 N+N (b)

~ '"Ol 3 o Total Chi a (r = 0.682)" (1) 0.1 - • > 20 ~ m Chi a (r = 0.48Qf >- .c '" 10- 20 ~m Chi a (r = 0.732)" a. 0 .II. 2-1 0 ~m Chi a (r = 0.603)' ~ ..Q • 0.2-2 Chi a (r = 0.656)' .c ~m U

0.01 0.01 0.1 1 SRP (c)

~ Ol Total Chi a (r = 0.390) '"3 • > 20 Chi a (r = 0.281 ) (1) 0.1 ~m '" 10-20 ~mChl a (r = 0.594)' >- .c a. .II. 2-10 ~ m Chi a (r = 0.517)' 0 ~ • 0.2-2 ~l m Chi a (r = 0.341) :c0 u

0.01 0.1 1 10 Silicic Acid (~M)

Figure 13. Correlations between (a) Chl a and N+N, (b) Chi a and SRP, and (c) Chi a and si licic acid. (*), significantly different from zero at the 99 % significance level. (+), significantly different from zero at the 95 % signjficance level.

41 2.3.3.2 Phycoerythrin

The latitudinal distribution of surface size-fractionated phycoerythrin concentrations are shown in Figures 14a and b. Highest concentrations of phycoerythrin were observed in waters located in the SPSG, PEQD and PNEC provinces and exhibited peaks both north and south of the equator at approximately 6° S and 6° N. The latitudinal distribution of surface phycoerythrin was dominated by picoplankton (0.2-2 JIlll) in all provinces with picoplankton phycoerythrin accounting for 75 - 95 % of the total phycoerythrin throughout the transect (Figure 14c). Surface picoplankton phycoerythrin concentrations ranged between 0.7 - 31 ng/l with the lowest levels of picoplankton phycoerythrin being observed in the NPTG (Figure 14a).

Nanoplankton (2-10 JIlll and 10-20 JIlll) phycoerythrin concentrations were markedly lower than those of picoplankton (Figures 14a and b). The highest phycoerythrin concentrations of nanoplankton (2-10 JIlll) also exhibited peaks both north and south of the equator at approximately 3° S and 4° N with concentrations reaching 1.73 ng/l.

Nanoplankton (2-10 JIlll) phycoerythrin accounted for < 10 % of the total phycoerythrin in each of the provinces (Figure 14c). The highest concentrations ofnanoplankton

(10-20 JIlll) phycoerythrin occurred in the SPSG and NPSG with concentrations reaching

1.16 ng/l. Nanoplankton (10-20 JIlll) phycoerythrin accounted for 15 % of the total phycoerythrin in the NPTG and < 5 % of the total phycoerythrin in each of the remaining provinces (Figure 14c).

42 Size - Fractionated Phycoerythrin 35.-__~SP~S~G ~-, ___P~E~Q~D ~-,~P~N~E~C~ ____~N~P~T~G~ __-. ~ >20 11m :::: 30 ~ 10 - 20 11 m g> 25 ~ 2-10Ilm ~ c: 20 ~ 0.2 -2 11m .'".r= ~ 15 8 10 >­ .r= 5 (L o

2 . 0~~------r------.-----'------, 1.8 (b) ~ >20 11m ::::- 1.6 ~ 10-20Ilm Cl E. 1.4 ~ 2- 10 11m c: 1.2 ..r='" 1.0 ~ 0.8 Ql o 0.6 u >- 0.4 .r= (L 0.2 0.0

-15 -10 -5 o 5 10 15 20 25 Latitude (Deg) 100.------~~------_, (c)

c: 80 .'".r= ~ 2l 60 u >­ .r= (L CI3 40 o I- ?J? 20

01----.... SPSG PEQO PNEC NPTG Province

Figure 14. Latitudinal distribution of surface size-fractionated phycoerythrin (a) >20 ~m , 10-20 ~m , 2-10 ~ and 0.2-2 ~m (b) >20 ~m , 10-20 ~m , and 2-10 ~m, and (c) Regionalized percent (%) total surface phycoerythrin. 43 Microplankton (>20 !J.IIl) phycoerythrin concentrations were also markedly lower than those of the picoplankton (Figures 14a and b). Microplankton phycoerythrin was present in surface waters with concentrations reaching 0.81 ngIl. Microplankton phycoerythrin concentrations accounted for 10 % of the total phycoerythrin in the NPTG province and

< 1 % of the total phycoerythrin in each of the remaining provinces (Figure 14c).

2.3.3.3 Adenosine 5 '- triphosphate (ATP)

The latitudinal distribution of surface total P-ATP concentrations are shown in Figure 15.

Total P-ATP concentrations throughout the transect ranged between 15 - 87 ngIl with increased variability in the PEQD and PNEC provinces. Highest levels of total P-A TP were observed in the equatorial PEQD and PNEC provinces between 5· S and 10· N.

Lowest levels of total P-ATP were observed in waters located in the oligotrophic SPSG and NPTG provinces.

The latitudinal distributions of surface size-fractionated P-A TP concentrations are shown in Figures 16a and b. Picoplankton P-ATP concentrations throughout the transect ranged between 3 - 20 ng/l, with highest concentrations occurring in the equatorial PEQD and

PNEC provinces (Figure 16b). Picoplankton P-ATP concentrations dominated the SPSG accounting for - 45 % of the total P-ATP in that province (Figure 16c). Picoplankton

P-ATP in the PEQD, PNEC and NPTG provinces accounted for 20 - 30 % of the total

P-ATP.

44 Adenosine 5'-triphosphate PEaD PNEC NPTG 100 SPSG

-+- P-ATP

80

:::::~ 60 OJ c ~ h 40 \ --- 20 ~ 1IV o -15 -1 0 -5 o 5 10 15 20 25 Latitude (Deg)

Figure 15 . Latitudinal distribution of surface total particulate ATP (P-ATP) and along the BEACH-BASH transect cruise.

45 Size - Fractionated ATP 60.-__~S~P~S~G ~-r __~P~E~Q~O~-,~P~N~E~C~ ____~N~P~TG~ ____~ (a) -- > 20 llm 50 -- 10 - 20 llm

::::- 40 Cl c::: ~ 30 (L I-- « 20

10 o~~~~~~~~tl~~~~~~~

30.------.------,-----.------~ (b) 25

:::- 20 Cl c::: ~ 15 (L ~ 10 •

5

o+-----.-----+-----,-----~----~----_r----_r----~ -15 -10 -5 o 5 10 15 20 25 Latitude (Oeg) 50.------~~------~------, (c) _ > 20 llm _ 10 - 20 llm 40 _ 2-101lm

(L ~ 30 (5'" I-- 20 'J<

10

o SPSG PEQO PNEC NPSG Province

Figure 16. Latitudinal distribution of surface size-fractionated particulate A TP (P-A TP) (a) >20 Iilll and 10-20 11m (b) 2-\ 0 11m and < 2 Iilll, and (c) Regionalized percent (%) total surface P-A TP. 46 Nanoplankton (2-10 ~ and 10-20 ~) P-ATP concentrations ranged between 3 - 27 ng/l and below detection limit « DL) -11 ng/l respectively (Figures 16aand b). The 2-10 ~ nanoplankton P-A TP concentrations showed greater variability than the 10-20 ~ nanoplankton P-ATP concentrations. P-ATP concentrations in the PEQD, PNEC and

NPTG provinces were all dominated by 2-10 ~ nanoplankton accounting for 35 - 45 % of the total P-ATP in those provinces (Figure 16c). However, the least dominant were the 10-20 ~ nanoplankton accounting for 5 - 15 % of the total P-ATP.

Microplankton (>20 ~) P-A TP concentrations throughout the transect ranged between

< DL - 53 ng/l, exhibiting the widest range out of all the size fractions (Figure 16a).

Microplankton P-ATP concentrations were lowest in the SPSG accounting for 10 % of the total P-A TP (Figure 16c). However, throughout the remainder of the transect microplankton P-ATP concentrations accounted for 25 - 30 % of the total P-ATP.

2.3.4 Plcoplankton community structure and toxonomic distribution

2.3.4.1 Flow Cytometry

The flow cytometric analysis of surface samples along the transect allowed for the enumeration of picoplankton (Prochlorococcus spp., Synechococcus spp., and picoeukaryotes) and heterotrophic bacteria. Patterns of abundance ofpicoplankton and heterotrophic bacteria in surface samples along the transect (Figure 17) aligned with some or all of the prominent features in N+N, SRP and Si(0In4-2 (Figure 18). Among all stations the estimates of surface heterotrophic bacteria abundances were highest ranging

4 4 l between 52 x 10 - 88 X 10 cells m1- (Figure 17a). High densities of heterotrophic

47 Plankton Abundances SPSG PEQO PN EC NPTG _ 100.---~~---,--~~~--~~~----~~~-----. "0 (a) ';; 90 E ;; 80 -a.'" 70 e ~ e 60 /

I* 50 +-----._----+_----._----+_----+_~--~ ~~~----_r~ ~~~ -15 -10 -5 o 5 10 15 20 25 ,,~ 20 .-~------._------._----~------_. (b) >< E 15

I':+- ____ ,-____ +_--~._----+_----+_----,_----,_--~ -15 -10 -5 o 5 10 15 20 25

,,~ 2.0 -y------:;:r.:------r::----,_------, ~ (c) i 1.5 -~ 1.0 I:: __ ------e ill -15 -10 -5 o 5 10 15 20 25

~ 1.0 ._------._------~----.------_. ~ (d) i 0.8 -~ 0.6 J!!'" g. 0.4 ~ § 0.2

a:: 0.0 +-----._----+_----,_----+_----+_----.------.----~ -15 -10 -5 o 5 10 15 20 25 Latitude (Oeg )

4 Figure 17. The latitudinal di stribution of surface plankton abundances ( # / ml x 10 ) (a) heterotrophic bacteri a (b) Prochlorococcus, (c) Synechococcus and (d) picoeukaryotes during the BEACH-BASH transect. 48 r = 0.58' r = 0.76' r = 0.39 ' 105 +---.---.-~~~

1o6 ,------,

~ u - 5 •• •• g L 10 ., .i. .. ~ ~ l·l·r·I'·'·:..J'·~· ••--': ":"'-'-,~~ , ' •• J ~ ~ 10' ;"- ., . .. . , o- •• ct

103 +-__.- __.-~ r _= T- ~0~.2~4

10 5~------, to .,. . ••• "-§~ 10' • • .. t> E .. • • o Vl . • • ""5=Q) Q) , 3 <::~ 10 6J •... "..• r = 0.52 ' r = 0.79' r = 0.32 10 2 +---.-~~~~~

105 ,------.

Vl 104 ~..-:>. .!.... •• ~ . E -. ... 3 ~ ~~.~---..~l ~ a; 103 a:o~ 10 2 +---.---.-~r _=~0~.3~3_1 r=O.73' r = 0.51 ' 10-3 10-2 10-1 1()O 10 1 10-2 10-1 10· 10 1 10-1 10· 101

Silicic Acid (I-lM)

Figure 18_ Correlations between the surface abundance of plankton groups (a) heterotrophic bacteria, (b) Prochlorococcus, (c) Synechococcus and (d) picoeukaryotes and nutrients along the BEACH-BASH transect. Left panels for +N, middle panels for SRP and right panels for Silicic acid. In some cases correlation coefficients (r) are indicated_

* Significantly different from zero at the 99 % significance level. + Significantly different from zero at the 95 % significance level. 49 bacteria were noted around the equator in the high-nutrient PEQD province. Lowest abundances of heterotrophic bacteria were observed in the low-nutrient SPSG and NPTG provinces. Plotting their abundance against all nutrient concentrations (Figure 18a) reveals a somewhat tighter relationship between SRP and abundance

2 (r = 0.76) than N+N and Si(OH)4- •

4 Most estimates of surface Prochlorococcus spp. abundance clustered within 1.5 x 10 -

4 l 10 X 10 cells mr\ with the highest density of15 x 104 cells mI· occurring in the PNEC province at approximately 6° N (Figure 17b). The lowest abundances of

Prochlorococcus spp. occurred just south of the equator in the high-nutrient PEQD province and in both the low-nutrient SPSG and NPTG provinces. Plotting their abundance against all nutrient concentrations (Figure 18b) reveals weak correlations with all nutrients. N+N and Si(OH)4-2 reveal a negative correlation and SRP reveals a slightly positive correlation.

Surface Synechococcus spp. abundances ranged between 0.04 x 104 - 1.88 x 104 cells mI-1 and exhibited peaks both north and south of the equator (Figure 17c). This distribution had a local minimum in Synechococcus spp. abundance centered about the equator with higher abundances of Synechococcus spp. to the north (6° N) and south

(5° S). The lowest Synechococcus spp. abundances were observed in the low-nutrient

NPTG province. The abundance of Synechococcus spp. along the entire transect correlated well with SRP and N+N (r = 0.79 and 0.52 respectively; Figure 18c).

so Surface picoeukaryote estimates among all stations were the lowest among the groups of

4 4 1 picoplankton studied ranging between 0.08 x 10 - 0.78 X 10 cells ml- (Figure 17d).

Picoeukaryote abundances exhibited a similar distribution to Prochlorococcus spp. with highest densities occurring in the PNEC province at approximately 6° N and lowest abundances occurring in both the low-nutrient SPSG and NPTG provinces. The abundance of picoeukaryotes correlated better with all nutrients compared to

Prochlorococcus spp. which showed weak correlations with nutrients. Picoeukaryotes correlated best with SRP and Si(OH)4-2 (r = 0.73 and 0.51 respectively; Figure 18d).

2.3.4.2 HPLC Pigments

The dominant pigments detected during the BEACH-BASH transect were monovinyl chlorophyll a (MVChl a), divinyl chlorophyll a (DVChl a), monovinyl plus divinyl chlorophyll b (TChl b), chlorophyll c-Iike pigments (TChl c), zeaxanthin (Zea), peridinin

(Per), diadinoxanthin (DDX), fucoxanthin (Fuco), 19'-butanoyloxyfucoxanthin (But),

19' -hexanoyloxyfucoxanthin (Hex) and a-carotene (a - Car). Concentrations of monovinyl chlorophyllide a (MVChld a), prasinoxanthin (Pras), alloxanthin (Allox), violaxanthin (Viola), lutein (Lut) and p-carotene (P - Car) were low and variable. or below the limit ofRP-HPLC quantification « DL). These results agree well with those found by Bidigare and Ondrusek (1996). Abbreviations and taxonomic affinities of photosynthetic pigments are shown in Table 2.

Distributions of total chlorophyll a (TChl a = MVChl a + DVChl a + MVChld a), total chlorophyll b (TChl b = MVChl b + DVChl b), and total chlorophyll c (TChl c) are

51 shown in Figure 19. The relationship between TChl a concentrations measured using the

TD-700 vs. TChl a concentrations measured by HPLC were also well correlated

(TD-700 = 0.89 (HPLC) + 0.013, ~ = 0.94,p < 0.001, n = 27; Figure 20). TChl a values ranged from 0.06 - 0.28 p.g/l with the highest concentrations occurring in the equatorial provinces (PEQD and PNEC). TChl a concentrations peaked north of the equator in the

PNEC province and were followed by an abrupt drop in concentration from 0.28 - 0.08 p.g/l. TChl b values ranged between 0.002 - 0.027 p.g/l with the highest concentrations occurring in the PEQD province and the lowest concentrations occurring in the NPTG province. TChl c values that ranged between 0.007 - 0.055 p.g/l. TChl c values followed a similar pattern to TChl a with highest concentrations occurring in the PEQD and PNEC equatorial provinces. TChl c concentrations also peaked north of the equator in the

PNEC province and were followed by an abrupt drop in concentration from 0.52 - 0.009 p.g/l. The lowest concentrations ofTChl c occurred in the NPTG.

The latitudinal distribution of surface TChl a pigment constituents MVChl a, DVChl a and MVChld a are shown in Figure 21a The chemotaXonomic marker MVChl a is representative of all phytoplankton (except Prochlorococcus spp.), DVChl a is representative of the prokaryotic pigment marker found in Prochlorococcus spp. and

MVChld a is representative of the degradation pigment found in senescent phytoplankton and fecal pellets. The latitudinal distribution ofMVChl a concentrations ranged between

0.029 - 0.146 p.g/l. The highest concentrations ofMVChl a occurred in the PEQD and

PNEC equatorial provinces and the lowest concentrations occurred in the oligotrophic

NPTG province. The latitudinal distribution ofDVChl a concentrations ranged between

52 HPLC Total Chlorophyll a, band c PEQD PNEC NPTG 0.30 SPSG 0.10 - TChia - TChlb 0.25 ~ TChic 0.08 ~ -=: ~ 0.20 ~ -=:~ <..l Ol 0.06 .c ::1. ~ () en 0.15 I- "0 .c c () 0.04 t'II I- .Q 0.10 .c () 0.02 I- 0.05 • 0.00 0.00 -10 -5 0 5 10 15 20 25 Latitude (Deg)

Figure 19. The latitudinal distribution of HPLC TChl a [MVChl a + DVChl a + MVChld a] , TChl b [MVChl b + DVChl b] and TChl c pigments.

53 0.30 • 0.25 • ~ ~ 0.20 ~ lIS :c (.) 0.15 • ~ • 0 • ....0 0 0.10 ~

0.05 y =0.89 x + 0.013, R2 =0.94 0.00 +------,---.------r----,.---,----I 0.00 0.05 0.10 0.15 0.20 0.25 0.30 HPLC TChl a ().tgll)

Figure 20. Chlorophyll a regression analysis. TChl a measured using the TD-700 versus TChl a measured by HPLC.

54 MVChl a, DVChl a and DVChld a

0.16 ,---..::..:...SPSG-=-:::.., __ --'--' PEQO==- _ -,---'--'PNEC-'==, ___ ----' NPTGc.::..-c....:::: ___----, (a) _ MVChla 0.14 -0 OVChl a 0.12 - MVChida -=::~ ~ 0.10

~ co 0.08 >. .c e0.. 0.06 .2 .c 0.04 () 0.02

0.00

-10 -5 o 5 10 15 20 25

90 ,----,------,----,------, (b) - MVChla 80 ~ OVChl a

<3 70 >o C/:I 60

:;:: 50 ~ :2 40 (ij ~ 30 :oR.o 20

10 +----+---r---~---4---._--_.--~ -10 -5 o 5 10 15 20 25 Latitude (Oeg)

Figure 21. The latitudinal distribution of (a) TChI a pigment constituents MVChl a, DVChl a and MVChld a, and (b) MVChl a and DVChl a as a proportion (%) ofTChl a [TChI a = MVChl a + DVChl a + MVChld aJ for the BEACH-BASH transect.

55 0.021- 0.1191l1Yl. The distribution of DVChl a closely follows the pattern oftlow cytometry Prochlorococcus spp. abundance (Figure 17b). DVChl a concentrations were highest in the PEQD and PNEC provinces and peaked in the PNEC province. Lowest

DVChl a concentrations occurred in the NPTG province. MVChld a concentrations were low and remained fairly constant throughout the transect with values of approximately

0.0021l1Yl. MVChld a concentrations show a single peak ofO.0351l1Yl at - 4.50 N latitude in the PEQD province.

The latitudinal distribution of the percent TChl a pigment markers MVChl a and

DVChl a are shown in Figure 21 b. MVChl a and DVChl a together accounted for

> 95 % of the TChl a throughout the transect. Percent total MVChl a and DVChl a both showed sinusoidal distributions, however they were inversely related to one another.

MVChl a accounted for approximately 50 - 60 % of the TChl a in the PEQD and NPTG provinces. DVChl a accounted for approximately 50 - 60 % of the TChl a in the SPSG and PNEC province suggesting that Prochlorococcus spp. was the dominant phytoplankton species in those provinces.

The latitudinal distribution of diagnostic pigments CyanoDP and ChromoDP as a proportion (%) of total diagnostic pigments [DP = Zea + DVChl a + TChl b + a-Car +

Hex + But + Fuco + TChl c] are shown in Figure 22. CyanoDP accounted for> 50 % total phytoplankton DP throughout the transect. CyanoDP accounted for approximately

70 - 80 % of total DP in the SPSG and NPTG provinces and 50 -70 % of total DP in the

PEQD and PNEC provinces. ChromoDP on the other hand accounted for < 50 % total

56 Cyanobacteria and Chromophyte Microalgae SPSG PEQD PNEC NPTG 90 - CyanoDP 80 - ChromoDP

70

c.. 60 0 III 50 -0 ~ ::§? 0 40

30

20

10 -10 -5 o 5 10 15 20 25 Latitude (Deg)

Figure 22. Diagnostic pigments CyanoDP and ChromoDP as a proportion (%) of the total diagnostic pigments [DP = Zea + DVChl a + TChl b + a-Car + Hex + But + Fuco + TChl c) for the BEACH-BASH transect cruise.

57 phytoplankton DP throughout the transect with 30 - 50 % occurring in the PEQD and

PNEC provinces and 20 - 30 % occurring in the SPSG and NPTG provinces. These results once again reinforce the dominance of picoplankton throughout the transect.

58 2.4 DISCUSSION

2.4.1 Nutrient dynll1llics

Nutrients in the oceans are circulated by biological and physical processes. In the euphotic zone of the surface ocean, nutrients are assimilated mainly by , including phytoplankton and bacteria. Biogenic particles are then carried below the mixed layer by the sinking of detritus and the feeding activities of animals, this particulate matter is then subject to remineralization which causes dissolved nutrients to be released back into seawater at depth. As a result, nutrient concentrations are generally low near the surface but high in deeper waters, except in Polar regions and/or regions of intensive upwelling like the equatorial Atlantic and Pacific which brings cold, nutrient­ rich seawater back to the surface. This study's principal transect, which exhibits a gradient in nutrients, was carried out during April-May 2005 (Figure 3). The transect began in the oligotrophic waters of the SPSG, passed through the nutrient enrich upwelled waters of the PEQD (-167° W), the chlorophyll enhanced waters of the PNEC, and ended in the oligotrophic waters of the NPTG coinciding with the HOT Station

ALOHA. The major nutrients (N+N, SRP and Si(OH)4'2) needed for phytoplankton growth increased in concentration in the surface waters of the equatorial Pacific between

5° S and the equator. Surface nutrient concentrations in this region exhibited rapid transitions associated with the strong westerly SEC and the region of equatorial divergence where nutrients are being upwelled.

Since phytoplankton require nutrients for their growth, nutrient stoichiometry and deviations from the Redfield ratio can limit phytoplankton primary production, and affect

59 phytoplankton biomass, species composition, and consequently food web dynamics. In the upper ocean, recycling ofN+N and phosphate are generally considered to be tightly coupled. Variations in nutrient supply, export or cellular incorporation can result in both local and regional changes in the N+N : SRP ratios. N+N: SRP stoichiometry across the transect never exceeded the Redfield molar ratio of 16N : IP and therefore imply a possible nitrogen limitation. The mean N+N : SRP ratio in the SPSG, PNEC and NPTG provinces was extremely low (mean = 0.08. n = 16). These results are similar to those found in the surface waters at Station ALOHA located in the NTPG province (Karl et m.,

2001b). However, measurements ofN+N and SRP by inorganic analysis alone could lead to erroneous conclusions about nutrient limitation (Downing, 1997; Karl et al.,

2001 b). Dissolved organic nutrients such as dissolved organic nitrogen (DON) and dissolved organic phosphorus (DOP) are also produced locally and consumed during and when included, have been shown to influence the N:P ratio

(Karl et aI., 2001b). Therefore, ifthe community was actually nitrogen limited, and no other source of fixed nitrogen were available (e.g. NH/ or DON) there would be a strong selection for the growth of diazotrophic microorganisms including the marine cyanobacteria Trichodesmium spp., the diatom endosymbiont Richelia and other unidentified N2-fixing bacteria capable of utilizing the nearly inexhaustible supply of nitrogen gas (Nz) in seawater and contributing to new production (Carpenter and

Ramons, 1991; Karl et al., 1992; 1997; Letelier and Karl, 1996; Capone et al., 1997; Zehr et al., 1998; 2001). In the PEQD province the N+N : SRP ratio associated with equatorial upwelling is closer to Redfield having a mean value of 7.6. With a N+N : SRP

60 ratio higher than that of oligotrophic regions one could expect a change in phytoplankton community structure, perhaps selecting against N2-fixing microorganisms.

While N+N and SRP are used to form soft tissues of organisms, Si(OH)4-2 is essential for the growth of the outer tests and of plankton such as diatoms, silicoflagellates and radiolarians. The dissolved Si(OH)4-2 : N+N ratio in the surface waters throughout the BEACH-BASH transect almost always exceeded the I: I ratio typical for diatom growth (Brzezinski, 1985). However, the ratio ofSi(OH)4-2 : N+N in equatorial upwelled source waters was less than the I: I ratio, which complements findings by

Wilkerson and Dugdale (1996) for this region. Assuming nutrients are constantly upwelled, an elevated Si(OH)4-2: N+N ratio > 1:1 would be indicative ofN-limitation, while a depressed ratio would be indicative ofSi(OH)4-2Iimitation (Conley and Malone.

1992). One explanation for the occurrence ofa Si(OH)4-2: N+N ratio < 1:1 in the equatorial region is a process called the "silicate pump" described in Dugdale et al.

(1995). The silicate pump works to export Si(OH)4-2 to deeper waters more efficiently than N+N. due to the faster regeneration rates ofN+N in surface waters relative to

Si(0H>4-2, and to the enhanced export ofSi(OH)4-2 that can occur when frustules are repackaged into fecal pellets, thus driving the system to Si-limitation. Si(OH)4-2 uptake by diatoms depends on light, temperature, the supply of nutrients and trace metals which brings us to another possible explanation. It has been suggested that phytoplankton growth in the equatorial Pacific is limited by iron (Fe) aVailability (Coale et al., 1996).

Studies have shown that in oceanic Fe addition experiments, controls were usually dominated by nanoplankton and picoplankton, whereas diatoms bloomed in Fe-enriched

61 incubations (Martin et aI., 1988; Hutchins, 1995; DiTullio et al., 1993). Under Fe limitation, Si(OHU·2 : N+N uptake ratios have been shown to be higher, suggesting diatoms stressed by a lack of iron should therefore deplete surface waters of silicic acid before nitrate, leading to a secondary silicic acid limitation of the phytoplankton community (Hutchins and Bruland, 1998). These results suggest that Fe-limitation may be an important factor driving the silicate pump (Hutchins and Bruland, 1998).

The dissolved inorganic Si(OH)4·2 : SRP ratio in the surface waters throughout the

BEACH-BASH transect almost never exceeded the Redfield ratio 16:1 (Redfield et al.,

1963), except in the extreme north at Station ALOHA with a value of 18. This

Si(OH)4"2 : SRP ratio of < 16: 1 would be indicative of Si-limitation of diatoms and therefore a dominance of either other large phytoplankton or small phytoplankton. The

Si(OH)4"2 : SRP ratio increased north along the transect, due to the increase in Si(OH)4"2 concentration relative to SRP concentrations.

2.4.2 Phytoplankton size structure

The equatorial Pacific. between 5° S - 0° in the PEQD province, despite the presence of excess nutrients in surface waters, is similar to other oligotrophic provinces in the tropical and subtropical open ocean in being dominated by picoplankton (0.2-2 J.1IO). Primary production was not measured on this cruise however, Chavez (1989) showed that cells

< I J.1IO accounted for about 50 - 60 % ofChl a standing stock and 50 % of 14C primary production over a wide expanse of the equatorial Pacific from 5° N to 5° S and from 90° to 180° W. Similarly, PeDa et a/. (1990) showed that the

62 fractions contributed 20 - 51 % and 76 - 87 % respectively to the total Chl a at 135° W, and Bidigare and Ondrusek (1996) showed that during the EqPac studies, 81- 92 % of

Chl a passed through 2-1= filters at 140° W. The fluorometric Chl a results from these studies are in close agreement with our data, in which picoplankton (0.2-2 1=) in the

PEQD province accounted for - 60 % ofChl a standing stock and between 50 - 60 % throughout the transect (14° S - 22° N).

The picoplankton contribution to TChl a increased in the PEQD province but was highest in the nutrient deplete waters of the PNEC province, away from equatorial divergence, and lowest in the oligotrophic waters of the SPSG and NPTG. Despite elevated macro­ nutrient concentration in the PEQD, phytoplankton chlorophyll concentrations are considerably lower than expected (Thomas, 1979). This inhibition of phytoplankton growth in the equatorial Pacific has been attributed to HNLC conditions (Minas el aJ.,

1986). The dominance of relatively small photosYnthetic cells in waters replete with N, P and Si has been taken as indirect evidence that primary production in this region of upwelling is limited by iron or some other trace nutrient (Chisholm, 1992). One explanation is that smaller cells are better able to use low dissolved Fe concentrations and thus are able to out-compete larger phytoplankton for Fe (Hudson and Morel, 1990). The supply of iron to the central and eastern equatorial Pacific is primarily from the

Equatorial Undercurrent (EUC), in which iron is drawn up to the euphotic zone by upwelling and the upward vertical mixing ofEUC waters (Gordon el aJ., 1997). The increase in TChl a north of the equator (- 6° N) in the PNEC province could be explained by advection of newly upwelled equatorial waters northward. Mechanisms for advection

63 include Ekman transport and the meridional currents associated with TrW's. Chl a standing stock would also be expected to increase if there was a growth-grazing imbalance or an input of the micro-nutrient iron via advection or atmospheric deposition.

Information on what governs phytoplankton community size structure in marine ecosystems is crucial to understanding what governs carbon flux in the ocean. Theories on phytoplankton species composition and size are based on nutrient concentrations and sinking properties (Margalef, 1978; Legendre and Le Fevre, 1989). Large cells, such as diatoms are observed to dominate on a biomass basis in high nutrient areas, while small cells, such as cyanobacteria, dominate in oligotrophic waters of the world's ocean.

However, as stated before, the HNLC region of the equatorial Pacific waters is dominated by smaller cells typical of oligotrophic regions. While it is difficult to say much about limiting factors based on correlations of size structure and nutrient concentrations, patterns did emerge. Correlation analysis ofTChl a and size-fractionated Chl a (Figures

13a, b and c) indicated that phytoplankton had a significant relationship with some of the nutrient concentrations. TChl a and size-fractionated Chl a concentrations showed a significant correlation to SRP concentrations. This correlation between TChl a, size­ fractionated Chl a and SRP could indicate SRP plays a role in controlling photosynthetic phytoplankton biomass. A positive correlation between Chl a and SRP, was observed throughout the transect; however, one might expect the opposite (a negative correlation), because phytoplankton growth would increase Chl a coincident with a drawdown in nutrients. However, there may be an imbalance in the time it takes to assimilate nutrients and create chlorophyll which may lead to this positive correlation. N+N and Si(OHk2

64 showed no significant correlation to TChi a, however, Si(OH)4·2 showed significant correlations with nanoplankton of the 2-10 J.I.lll and 10-20 J.I.lll size fraction. This correlation between nanoplankton Chi a and Si(OH)4·2 may be an indication that

Si(OH)4·2 is controlling diatom abundance. Previous studies done by JGOFS, and IronEx

II, revealed an equatorial diatom community dominated by small « 10 J.I.lll), rapidly growing pennate diatoms, some of which were being grazed on by microzooplankton and therefore becoming part of the microbial food web (Latasa et aI., 1997; Landry et aZ.,

2000a; 2000b; 2003). N+N showed no significant correlation with TChi a or size­ fractionated Chi a, these results were similar to those observed by Chavez (1989).

Phycoerythrin is an accessory pigment which can be used as a biomarker for cyanobacteria (Synechococcus spp. and Trichodesmium spp.) and rhodophytes (Stewart and Farmer, 1984). The cyanobacteria Synechococcus spp. and Trichodesmium spp. that contain phycoerythrin have been found to be important contributors to primary production in the world's oceans (Waterbury et aZ., 1979; Letelier and Karl, 1996).

Surface phycoerythrin concentrations throughout the BEACH-BASH transect were overwhelmingly dominated by picoplankton (0.2-2 J.I.lll) accounting for 75 - 95 % of the total phycoerythrin. This dominance of picoplankton phycoerythrin throughout the transect closely matched Synechococcus spp. abundance and exhibit a significant correlation (r = 0.83). Neveux et al. (1999) also found significant correlations between phycoerythrin and Synechococcus spp. in equatorial and south Pacific waters. Their study also revealed Synechococcus spp. in oligotrophic and eutrophic waters obtained a different phycoerythrin excitation spectra indicating that different Synechococcus spp.

65 may have different ecological requirements. While 75 % of the total phycoerythrin in the

NPTG was attributed to picoplankton, in this province - 20 % of the total phycoerythrin was due to the contribution of organisms in the >20 J.Un and 10-20 J.Un size fraction.

Phycoerythrin associated with these size fractions is probably attributed to large cyanobacterial diazotrophs, which include filamentous and colonial Trichodesmium and diatoms containing endosymbiotic Richelia (Zehr et at., 2001; Karl et at., 2002; Dore et at., 2002).

There are a number of phytoplankton pigment markers like chlorophyll, carotenoids and phycoerythrin which are often used for estimating the presence of individual groups of microorganisms. ATP measurements, on the other hand, can and have been widely used to estimate total microbial biomass (Holm-Hansen and Booth, 1966; Karl, 1980; Karl and

Dobbs, 1998). This holds true, because without exception, all living organisms contain

ATP. Measurements of ATP concentrations are therefore capable of capturing organisms within a community that display a wide variety of nutritional modes such as photoautotrophy, mixotrophy and/or heterotrophy. P-ATP concentrations have been measured extensively in the NPTG province at Station ALOHA (Winn et at., 1995;

Bj6rkman and Karl, 2005). However, to my knowledge P-A TP concentrations have rarely, if ever been measured in the central equatorial Pacific. Throughout the BEACH­

BASH transect P-ATP concentrations were highly variable and exhibited a rapid transition starting at the Equatorial Front (-5· S). This variability is most likely attributed to physical processes such as Kelvin waves which modulate the depth of the nutrients in the EUC and by TIW's which are a common phenomenon in the central equatorial

66 Pacific and result from the cyclonic shear between the EVC and the SEC (Luther and

Johnson, 1990).

Size-fractionated P-ATP measurements also exhibited regional variability. In the SPSG, picoplankton dominated total biomass, accounting for - 45 % of the total P-ATP, consistent with increased abundance ofheterotropbic bacteria in this province (Figure

17a). Since size-fractionated Chl a and phycoerythrin data suggests the dominance of photosynthetic picoplankton throughout the transect one would expect similar trends in

P-ATP. However, in PEQD, PNEC and NPTG provinces P-ATP concentrations are mainly dominated by organisms in the 2-10 f.UIl and >20 f.UIl size fraction, suggesting that larger heterotropbic organisms, not accounted for in Chl a measurements may be playing an important role in size structure and therefore food web dynamics. Hagstr6m et al.

(1988) suggested that although microplankton can effectively graze picoplankton populations, it is the nanoplankton that are responsible for the majority of picoplankton clearance. VIMS et aI. (1995) showed that during the EqPac studies the heterotrophic protistan communities were dominated by nanoplankton , such as dinoflagellates, , as well as naked flagellates such as amoebae.

Brown et aI. (2003) also found that nearly all of the larger heterotrophic protistan ~ 8 f.UIl grazers were dinoflagellates.

2.4.3 Community structure and taxonomic composition

Picoplankton are now widely recognized as dominating biomass, production and metabolic activity in the open oceans (Williams, 1981; Azam et al., 1983; Li et al., 1983;

67 Takahashi and Bienfang, 1983; Murphy and Haugen, 1985; Iturriaga and Mitchel, 1986).

The composition of the microbial community in the equatorial Pacific has been characterized as being dominated by small cells and deficient in large bloom-forming diatoms (Le Bouteiller et al., 1992; Mackey et aI., 2002; Landry and Kirchman, 2002;

Brown et aZ., 2003). Flow cytometry has been instrwnental in demonstrating the structure of the picoplankton community in the equatorial Pacific. It has shown that the equatorial Pacific is dominated by heterotrophic bacteria, two populations of cyanobacteria (Prochlorococcus spp. and Synechococcus spp.) and a mixed population of picoeukaryotes. However, less well understood is the community iesponse to varying physical forces and the relationship between microbial abundances and nutrient concentrations.

The picoplankton of the tropical Pacific have been described as the background over which the dynamic of the larger organisms are imprinted (Landry and Kirchman, 2002).

The abundance of heterotrophic bacteria during the BEACH-BASH transect was highest in the PEQD province, just south of the equator in nutrient replete waters and generally decreased from equatorial to subtropical waters. These results are similar to those found by Landry and Kirchman (2002) which compared heterotrophic abundances from three different equatorial Pacific cruises (EqPac TIOII, EqPac TI007 and EBENE). Landry and Kirchman (2002) found that among the three cruises heterotrophic abundances were tightly constrained south of the equator and among the three cruises no systematic differences were observed. In contrast, north of the equator heterotrophic bacterial abundance differed greatly among the three cruises. Our data most closely resembled the

68 La Nifia EqPac cross equatorial transect (120 N - 120 S, 1400 W) cruise ITOll with relatively high abundances of heterotrophic bacteria to the north. Patterns in the abundance of heterotrophic bacteria and picopiankton in surface samples along the

o2 transect aligned with some of the prominent features in N+N, SRP and Si(OH)4

(Figure 18).

Throughout the BEACH-BASH transect Prochlorococcus spp. was by far the most nwnerica1ly dominant photosynthetic picoplankton. Prochlorococcus spp. exhibited a maximwn abundance occurring at 60 N, and a minimwn abundance in the SPSG, NPTG and in the region of equatorial upwelling in the PEQD. Thus, Prochlorococcus spp. probably accounted for most of the Chi a observed in the smallest (0.2-2 ~) size fraction. While previous studies also found Prochlorococcus spp. the most dominant autotrophic group, abundances observed in this study were a factor offour lower (Binder et 01., 1996; Blanchot et al., 200 I; Mackey et 01., 2002; Landry and Kirchman, 2002;

Brown et 01., 2003). Together, these past studies showed an average surface

Prochlorococcus spp. abundance of approximately 20 x 104 cells mrl whereas, average abundances during this study were approximately 5 x 104 cells ml°l. These differences in abundances could be attributed to seasonal or climatic variations (Landry et al., 1996).

Over the 15 years of records of Prochlorococcus spp. abundances in the surface waters of

Station ALOHA (Figure 23a) there were periods of relatively high (1995-2000) and relatively low Prochlorococcus sppo abundances (2000-2005). Karl et 01. (2001a) hypothesized that at Station ALOHA a possible "domain shift" occurred leading to an ecosystem dominated by in response to climatic variations. Perhaps these

69 40.------~------~----~ (a)

x E 30

a. a. (J) 20

~ '"tl 8 Qe 10 1i Qe

88 89 90 91 92 93 94 95 96 97 98 99 00 01 02 03 04 05 06 07 Year 30 .------, (b) 20

10

o o (/) -10

-20

-30 _ SOl --- 5 month weighted mean ~O+--,_,--r-,-_r_.--,_,__r_,--,__r_._,r_.__._,--~ 88 89 90 91 92 93 94 95 96 97 98 99 00 01 02 03 04 05 06 07 Year

Figure 23. (a) Temporal di stribution of Prochlorococcus spp. abundance at Station ALOHA between 1991-2005 (integrated from 0-10 m) . (b) Southern Oscillation Index (SOl) between 1988-2007.

70 differences observed at Station ALOHA can be seen throughout the entire Pacific basin and are a reflection of climate forcing and the Southern Oscillation Index (SOl) (Figure

23b).

While it is difficult to say much about limiting factors based on correlations of plankton abundan~s and nutrient concentrations, patterns did emerge. Prochlorococcus spp. showed a slightly negative correlation with N+N and no correlation with SRP and

Si(0H)4·2. These results were similar to those found in the Atlantic by Cavender-Bares et al. (2001). Cavender-Bares et al. (2001) found that temperature was an important regulating factor for Prochlorococcus spp.; however, no relationship between temperature and Prochlorococcus spp. abundance was observed in this study (data not shown). Recent laboratory studies have shown that cultures of Prochlorococcus spp. cannot utilize nitrate for growth (Rippka et al., 2000; Moore et al., 2002), this could possibly explain the low abundances of Prochlorococcus spp. found in the N+N replete waters of the equatorial Pacific. Another possible explanation is perhaps

Prochlorococcus spp. are being out-competed when nutrients are abundant because they are either unavailable to them (e.g., nitrate) or their maximum growth rate is not as fast as that of other species such as Synechococcus spp. (Moore et al., 1995).

Surface Synechococcus spp. abundances throughout the BEACH-BASH transect

exhibited a local minimum in abundance centered about the equator (- 2° N) with higher abundance to the north (6° N) and to the south (S° S) in the N+N replete waters. The

lowest abundances of Synechococcus spp. were observed in the oligotrophic waters of the 71 SPSG and NPTG provinces. These results closely resemble those found during the

EqPac cross equatorial transect (120 N - 120 S. 1400 W) cruises IT007 and IT011 in

February-March 1992 and August-September 1992, respectively (Landry et al., 1996).

During these cruises, Landry et aI. (1996) found that climatic events such as El Nillo

(IT007) generally increased Synechococcus spp. abundances compared to La Nifta

(ITO 1 I) conditions. The BEACH-BASH transect took place during normal conditions a few months after a slight warming event and exhibited similar surface distributions to the

La Nifta EqPac cruise IT011. Surface Synechococcus spp. abundances along the transect

in contrast to Prochlorococcus spp. abundances correlated well with surface SRP and

N+N concentrations. Tbis may be because Synechococcus spp. can utilize nitrate as a nitrogen source (Waterbury et aI., 1986) and therefore possibly out compete

Prochlorococcus spp. which showed lower abundances in the N+N replete waters of the equatorial Pacific.

Picoeukaryote abundances were the lowest among the groups of photosynthetic picoplankton studied, exhibiting a similar distribution to Prochlorococcus spp. through out the transect. These results are similar to those found by Landry et al. (1996) where high surface abundances of picoeukaryotes occurred on the northern side of the transect during the EqPac cruise IT007. In contrast to previous comparisons in which

Synechococcus spp. and heterotrophic bacteria abundances closely matched the La Nifta

EqPac cruise IT01l, picoeukaryotes abundances matched the El Nillo EqPac cruise

IT007. Since the timing of the BEACH-BASH cruise took place during a normal ENSO phase, perhaps the picoplankton community response to environmental forcing reflects 72 ENSO phases. Picoeukaryote abundance correlated best with SRP and Si(OH)4-2 concentrations.

The distributions of chlorophyll and carotenoid pigments throughout the BEACH-BASH transect allowed us to examine the spatial variations in phytoplankton pigments, and taxonomic composition in the Pacific. TChl b [MVChl b + DVChl b] was a relatively abundant accessory chlorophyll in the equatorial and tropical Pacific. Potential sources for MVChl b detected in surface waters are not always clear (Neveux et al., 2003).

However, groups that typically contain MVChl b include chlorophytes and prasinophytes and DVChl b is a light harvesting pigment found exclusively in Prochlorococcus spp.

(Bidigare and Ondrusek, 1996). Concentrations of carotenoids relative to chlorophytes and prasinophytes were low and variable throughout the transect, these results matched those ofBidigare and Ondrusek (1996). This along with the work ofPartenskey et aZ.

(1993) which suggests that some strains of Prochlorococcus spp. are indeed able to synthesize MVChl b and DVChl b at high growth irradiances lead them and others

(Landry et al., 2000b and 2003) to conclude that most of the MVChl b and DVChl b at the surface in the enriched equatorial area probably belonged to Prochlorococcus spp.

TChl c concentrations were also an abundant accessory chlorophyll with highest concentrations occurring in the PEQD. Some of the primary organisms that contain

TChl c include prymnesiophytes, pelagophytes and diatoms (Jeffrey et al., 1997).

The ability to quantify reliably MVChl a, a pigment found in all phytoplankton (except

Prochlorococcus spp.) and DVChl a, a pigment marker found in Prochlorococcus spp.

73 allows us to look at the distribution of this prochlorophyte and to assess its contribution to the total plankton biomass «DVChl a I (MVChl a +DVChl a +MVChld a))*100 %).

Overall, DVChl a accounted for 50 - 60 % of the TChl a throughout the BEACH-BASH transect and agreed well with our flow cytometry data. This contribution ofDVChl a to

TChl a reinforces the importance of these picoplankton (i.e. Prochlorococcus spp.) to the phytobiomass over extensive oceanic regions. The latitudinal distribution shows a decrease in the contribution of DVChl a to TChl a in the PEQD province. Since

DVChl a, is a major pigment marker for Prochlorococcus spp. we could potentially attribute this drop in DVChl a to a possible nutrient limitation, such as Fe, or as

mentioned earlier to the enhancement ofN+N in surface waters. This is confirmed by

looking at the contribution of diagnostic pigments to chromophyte microalgae

(ChromoDP; i.e. prymnesiophytes, pelagophytes and diatoms) and to cyanobacteria and

prochlorophytes (CyanoDP; i.e. Prochlorococcus spp. and Synechococcus spp.).

Throughout the transect the concentrations ofCyanoDP dominated. These results

complement our size-fractionated and flow cytometry data obtained during this study.

However, in the PEQD province the proportion of CyanoDP decreased as ChromoDP

increased, suggesting that prymnesiophytes. pelagophytes and diatom abundances

increased in the nutrient rich waters of the equatorial Pacific although they did not

dominate. These high concentrations ofChromoDP (Hex, But, Fuco and TChl c) are

consistent with previous studies which report relatively high abundances of some if not

all of the these diagnostic pigments in the equatorial Pacific (Chavez et aZ., 1990;

Bidigare and Ondrusek, 1996; Mackey et al., 1998; Higgans and Mackey, 2000; Mackey

et al., 2002; Brown et aZ., 2003). 74 2.4.4 Implications for the biological pump

Currently, the conceptual view of carbon export from the oceanic surface layer is controlled by the biological transfo~ons that occur within the food web. Small photosynthetic cyanobacteria and heterotrophic bacteria which are thought to dominate oligotrophic open ocean ecosystems are believed to contribute relatively little to carbon export from the surface layers to the deep because of their small sizes, slow sinking rates and there efficient utilization of recycled organic carbon in the microbial loop (Legendre and Le Fevre, 1989; Michaels and Silver, 1998). Larger, faster growing phytoplankton such as diatoms are thought to dominate during increased nutrient availability and are believed to be more efficient in transferring carbon from the surface layers to the deep because of their large sizes, rapid sinking rates and their availability for predation by higher trophic levels (i.e. and fish larvae; Legendre and Le Fevre, 1989;

Michaels and Silver, 1998) and for export through sedimentation (Ki0rboe, 1993).

However, here I have shown that in the equatorial Pacific, despite elevated macronutrient concentrations, surface waters are similar to oligotrophic provinces in being dominsted by small picoplankton (0.2-2 IIID). This therefore suggests that a strong and stable microbial food web exists in the equatorial Pacific, which is thought to derive from a

close coupling of growth, grazing and reminera1ization (Landry et al., 1991; 2000a;

2000b; Landry and Kirchman, 2002). Estimates of new production using stable nitrogen

isotopes have also found low f -ratios, which indicated highly efficient recycling of

nitrogen (Murray et aI., 1989; Price et al., 1991; Pella et al., 1992; Dugdale et al., 1992;

McCarthy et al., 1996).

15 A recent study by Richardson and Jackson (2007) which used an inverse model approach to trace carbon fluxes through food webs suggests that the contributions of picoplankton to carbon flux from the surface waters to the deep have been overlooked and that in fact picoplankton contribute to export in the form of aggregates, which create larger particles that can sink faster and contribute to export either directly though sinking or indirectly through consumption. The works of Smith et al. (1996) during the EqPac studies have reported abyssal accumulations of phytodetritus in the central equatorial Pacific. Another recent study by Salihoglu and Hofinann (2007) which used a one-dimensional multi­ component lower trophic level ecosystem model that included detailed algal physiology was used to investigate the response of phytoplankton communities and carbon production and export to variations in physical and biochemical processes in the equatorial Pacific. Their results show that high-frequency variability in vertical advection and temperature is an important mechanism driving carbon export. For example, times of increased stratification can result in decreased iron concentrations and therefore reduced vertical velocities which would lead to decreased carbon export. From these studies it is clear that both biological and physical processes contribute to the efficiency of the biological carbon pump and the resulting magnitude of carbon export to the deep sea.

76 CHAPTER 3. CONCLUSION

Investigations of nutrient inventory and dynamics, phytoplankton size structure, community composition and succession, and rates of primary productivity are key variables in marine ecological research. These processes depend on a variety of time and space scales and together determine the ocean's ability to sequester carbon dioxide from the atmosphere. The BEACH-BASH transect was a single cruise, or "snapshot in time", focusing on phytoplankton size structure and community composition across biogeochemical gradients in the Pacific Ocean. The grouping of stations into biogeochemical provinces as defined by Longhurst (1998), allowed me to compare data over an extensive spatial scale.

The observations made during this and previous studies have contributed to the knowledge of phytoplankton dynamics in the equatorial Pacific Ocean. Observed variations found during this and other studies were most likely caused by a combination of physical (Kelvin waves, TIWs and advection), chemical (iron limitation). biological

(growth-grazing imbalances) and climatic (EI Niiio or La Nifta ENSO phase) processes.

However, it is clear that the finer details of the observed variability in phytoplankton size-structure, abundances and community composition will only be resolved when more samples are collected with simultaneous measurements on the appropriate time and space scales. For example, moorings with sensors that can produce high resolution measurements on chemical and biological properties such as nitrate, oxygen and fluorescence would complement the Tropical Atmosphere Ocean (TAO) arrays which 77 measure physical and atmospheric properties throughout the equatorial Pacific.

Comparisons to Station ALOHA throughout this study bring about the importance of time-series measurements. Repeated oceanographic measurements are crucial to understanding natural habitat variability as well as climate influences. Therefore, conducting repeat transects in the future would facilitate a better understanding on the controlling factors of phytoplankton dynamics over wide spatial scales in the Pacific

Ocean and significantly improve our understanding of variability in the system. Lastly, the incorporation of data into models will aid in understanding and predicting the complex interaction between biological and chemical processes.

78 REFERENCES

Azam F, Fencbel T, Field JG, Gray JS, Meyer-Reil LA, Thingstad F. 1983. The ecological role of water-column microbes in the sea Marine Ecology Progress Series 10: 257-63.

Barber RT, Ryther JH. 1969. Organic Chelators: Factors affecting primary production in the Cromwell Current upwelling. Journal of Experimental Marine Biology and Ecology 3: 191-199.

Barber RT, Chavez FP. 1986. Ocean variability in relation to living resources during the 1982-83 El Nillo. Nature 319: 279-285.

Barlow RO, Mantoura RFC, Oough MA, Fileman TW. 1993. Pigment signatures of the phytoplankton composition in the northeastern Atlantic during the 1990 . Deep-Sea Research n 40: 459-477.

Barlow RO, Aiken J, Holligan PM, Cummings DO, Maritorena S, Hooker S. 2002. Phytoplankton pigment and absorption characteristics along meridional transects in the Atlantic Ocean. Deep-Sea Research I 49: 637-660.

Barlow RO, Aiken J, Moore OF, Holligan PM, Lavender S. 2004. Pigment adaptations in surface phytoplankton along the eastern boundary of the Atlantic Ocean. Marine Ecology Progress Series 281: 13-26.

Bidigare RR, Ondrusek ME. 1996. Spatial and temporal variability of phytoplankton pigment distributions is the central equatorial Pacific Ocean. Deep-Sea Research n 43: 809-833.

Bidigare RR, Van Heukelem L, Trees CC. 2005. Analysis of algal pigments by high performance liquid chromatography. Algal Culturing Techniques. Academic Press. p 327-345.

Binder BJ, Chisholm SW, Olson RJ, Frankel SL, Worden AZ. 1996. Dynamics of picophytoplankton, ultraphytoplankton and bacteria in the central equatorial Pacific. Deep sea Research n 43: 907-931.

Bj6rkman KM, Karl DM. 2005. Presence of dissolved nucleotides in the North Pacific Subtropical Oyre and their role in cycling dissolved organic phosphorus. Aquatic 39: 193-203.

Blanchot J, Andre J-M, Navarette C, Neveux J, Radenac M-H. 2001. Picophytoplankon in the equatorial Pacific: vertical distributions in the warm pool and in the high nutrient low chlorophyll conditions. Deep-Sea Research I 48: 297-314.

79 Blanco 1M, Echevarria F, Garcia CM. 1994. Dealing with size-spectra: some conceptual and mathematical problems. Scientia Marina 58:17-29.

Brown SL, Landry MR, Neveux J, Dupouy C. 2003. Microbial community abundance and biomass along a 180' transect in the equatorial Pacific during an El Niilo­ Southern Oscillation cold phase. Journal of Geophysical Research 108(CI2): 8139, doi: 10.10291200IJC000817.

Brzezinski MA. 1985. The Si:C:N ratio of marine diatoms: interspecific variability and the effect of some environmental variables. Journal ofPhycology 21:347-357

Campbell L, Vaulot D. 1993. Photosynthetic picoplankton community structure in the subtropical North Pacific Ocean near Hawaii (Station ALOHA). Deep-Sea Research 40: 2043-2060.

Campbell L, Nolla HA, Vaulot D. 1994. The importance of Prochlorococcus to community structore in the central North Pacific Ocean. Linmology and Oceanography 39: 954-961.

Capone DG, Zehr JP, Paerl HW, BergmanB, Carpenter EJ. 1997. Trichodesmium, a globally significant marine cyanobacterium. Science 276: 1221-1229.

Carpenter EJ, Romans K. 1991. Major role of the cyanobacterium Trichodesmium in nutrient cycling in the North Atlantic Ocean. Science 254: 1356-1358.

Cavender-Bares KK, Karl DM, Chisholm SW. 2001. Nutrient gradients in the western North Atlantic Ocean: Relationship to microbial community structure and comparison to patterns in the Pacific Ocean. Deep-Sea Research I 48: 2373-2395.

Chavez FP. 1989. Size distribution of phytoplankton in the central and eastern tropical Pacific. Global Biogeochemical Cycles 3: 27-35.

Chavez FP, Barber RT. 1987. An estimate of new production in the equatorial Pacific. Deep Sea Research 34: 1229-1243.

Chavez FP, Buck KR, Barber RT. 1990. Phytoplankton taxa in relation to primary production in the Equatorial Pacific. Deep-Sea Research 37: 1733-1752.

Chavez FP, Buck KR, Coale KH, Martin JH, DiTullio GR, Welschmeyer NA, Jacobson AC, Barber RT. 1991. Growth rates, grazing, sinking and iron limitation of equatorial Pacific phytoplankton. Linmology and Oceanography 36: 1818-1833.

Chavez FP, Buck KR, Bidigare RR, Karl DM, Hebel D, Latasa M, Campbell L. 1995. On the chlorophyll a retention properties of glass-fiber GFIF filters. Linmology and Oceanography 40: 428-433.

80 Chisholm SW. 1992. Phytoplankton size. Primary Productivity and Biogeochemical Cycles in the Sea. In: Falkowski PG, Woodhead AD, editors. Plenum Press, New York.

Chisholm SW, Olson RJ, Zettler ER. Goericke R. Waterbury JB, Welscbmeyer NA. 1988. A novel free-living prochlorophyte abundant in the oceanic euphotic zone. Nature 334: 340-343.

Chisholm SW, Frankel SL, Goericke R. Olson RJ, Palenik B, Waterbury JB, West­ Johnsrud L, Zettler ER. 1992. Prochlorococcus Marinus nov. gen. nov. sp.: an ocyphototrophic marine containing divinyl chlorophyll a and b. Archives of 157: 297-300.

Coale KH, Johnson KS, Fitzwater SE, Gordon RM, Tanner S, Chavez FP, Ferioli L, Sakamoto C, Rogers P, Millero F, Steinberg P, Nightingale P, Cooper D, CochJan WP, Landry MR, Constantinou J, Rollwagen G, Trasvina A, Kudela R. 1996. A massive phytoplankton bloom induced by an ecosystem-scale experiment in the equatorial Pacific Ocean. Nature 383: 495-501.

Conley OJ, Malone TC. 1992. Annual cycle of dissolved silicate in Chesapeake Bay: implications for the production and fate of phytoplankton biomass. Marine Ecology Progress Series 81: 121-128.

Cox RD. 1980. Determination of nitrate at the parts per billion level by chemiluminescence. Analytical Chemistry 52: 332-335.

Cullen JJ. 1991. Hypotheses to Explain High-Nutrient Conditions in the Open Sea. Limnology and Oceanography 36: 1578-1599

Delcroix T, Picaut J. 1998. Zonal displacement of the westero equatorial Pacific "fresh pool". Journal of Geophysical Research CI03: 1087-1098.

Dickey TD. 1991. The emergence of concurrent high-resolution physical and bio-optical measurements in the upper ocean and their applications. Reviews of geophysics 29: 383-413.

Dickson ML, Wheeler PA. 1993. Chlorophyll a concentrations in the North Pacific: Does a latitudinaJ gradient exist? Limnology and Oceanography 38: 1813-1818.

DiTullio GR. Hutchins DA, Bruland KW. 1993. Interaction of iron and major nutrients controls phytoplankton growth and species composition in the tropical North Pacific Ocean. Limnology and Oceanography 38: 495-506.

81 Dore JE, Brum JR, Tupas LM, Karl DM. 2002. Seasonal and Interannual Variability in Sources of Nitrogen Supporting Export in the Oligotrophic Subtropical North Pacific Ocean. Limnology and Oceanography 47: 1595-1607.

Downing JA. 1997. Marine Nitrogen: phosphorus stoichiometry and the global N:P cycle. Biogeochemistry 37: 237-252.

Dugdale RC, Goering JJ. 1967. Uptake of new and regenerated forms of nitrogen in primary productivity. Limnology and Oceanography 12: 196-206.

Dugdale RC, Wilkerson FP, Barber RT, Chavez FP. 1992. E....rimating new production in the equatorial Pacific Ocean at 150' W. Journal of Geophysical Research 97: 681- 686.

Dugdale RC, Wilkerson FP, Minas ill. 1995. The role ofa silicate pump in driving new production. Deep-Sea Research I 42: 697-719.

Falkowski PG. 1994. The role of phytoplankton photosynthesis in global biogeochemical cycles. Photosyn Res 39: 235-258.

Feeley RA, Gammon RH, Taft BA, Pullen PE, Waterman LS, Conway TJ, Gendron JF. 1987. Distribution of chemical tracers in the eastern equatorial Pacific during and after the 1982-83 El Nifio-Southern oscillation event. Geophysical Research Letters 21: 277-280.

Fernandez E, Marail6n E, Moran XAG, Serret P. 2003. Potential causes for the unequal contribution of picoplankton to total biomass and productivity in oligotrophic waters. Marine Ecology Progress Series 254: 101-109.

Fiedler PC, Philbrick V, Chavez FP. 1991. Ocean upwelling productivity in the eastern tropical Pacific. Limnology and Oceanography 36: 1835-1850.

Garside C. 1982. A chemiluminescent technique for the determination of nanomolar concentrations of nitrate and nitrite in sea-water. Marine Chemistry 11: 159-167.

Gibb SW, Barlow RG, Cummings 00, Rees NW, Trees CC, Holligan P, Sugget D. 2000. Surface phytoplankton pigment distribution in the Atlantic Ocean: An assessment ofbasinscale variability between 50' N and 50' S. Progress in Oceanography 45 (3-4): 339-368.

Gordon RM, Coale KH, Johnson KS. 1997. Iron distributions in the equatorial Pacific: implications for new production. Limnology and Oceanography 42: 419-431.

82 Hagstr6m A, Azam F, Andersson A, Wilmer J, Rassoulzadegan F. 1988. Microbialloop in an oligotrophic pelagic ecosystem: possible roles of cyanobacteria and nanoflagellates in the organic fluxes. Marine Ecology Progress Series. 49: 171- 178.

Higgins HW, Mackey OJ. 2000. Algal class abundances, estimated from chlorophyll and carotenoid pigments, in the western Equatorial Pacific under El Nino and non-El Nino conditions. Deep-Sea Research I 47: 1461-1483.

Holm-Hansen 0, Booth CR. 1966. The measurement of adenosine triphosphate in the ocean and its ecological significance. Limnology and Oceanography 11: 510-519.

Holm-Hansen 0, Lorenzen CJ, Holmes RW, Strickland IDH. 1965. Fluorometric determination of chlorophyll. Journal du Conseil Permanent International pour l'Exploration de la Mer 30: 3-15.

Hudson RJM, Morel FMM. 1990. Iron transport in marine phytoplankton: kinetics of cellular and medium coordination reactions. Limnology and Oceanography 35: 1002-1020.

Hutchins DA. 1995. Iron and the marine phytoplankton community. In: Chapman D, Round F, editors. Progress in Phyc%gica/ Research Vol. II. Biopress, Bristol. 1-49

Hutchins DA, Bruland KW. 1998. Iron-limited diatom growth and Si:N uptake ratios in a coastal upwelling regime. Nature 393: 561-564.

Inoue HY, Ishii M, Matsueda H, Ahoyama M. 1996. Changes in longitudinal distribution of the partial pressure of C(h (pe02) in the central and western equatorial Pacific, west of 160· W. Geophysical Research Letters 23: 1781-1784. lturriaga R, Mitchell BO. 1986. Chroococcoid cyanobacteria: A significant component in the food web dynamic of the open ocean. Marine Ecology Progress Series 28: 291-297.

Jeffrey SW, Mantoura RFC, Wright SW, editors. 1997. Phytoplankton pigments in Oceanography. Monographs on Oceanographic Methodology. UNESCO publishing, Paris, 661.

Johnson PW, Sieburth J McN. 1979. Chroococcoid cyanobacteria in the sea: a ubiquitous and diverse phototrophic biomass. Limnology and Oceanography 24: 928-935.

Karl DM. 1980. Cellular nucleotide measurements and applications in microbial ecology. Microbiological Reviews 44: 739-96.

83 Karl DM. 1993. Total Microbial biomass estimation derived from the measurement of particulate adenosine-5' -triphosphate. In Kemp PF, BF, Sherr EB, Cole JJ, editors. Current methods in aquatic microbial ecology. Lewis Publishers, Boca Raton, Florida. p 359-368.

Karl DM. 1999. A sea of change: biogeochemical variability in the North Pacific subtropical gyre. Ecosystems 2: 181-214.

Karl DM. 2002. Nutrient dynamics in the deep blue sea. Trends in Microbiology, Vol. 10,9: 410-418.

Karl DM, Bidigare RR, Letelier RM. 2001a. Long-term changes in plankton community structure and productivity in the North Pacific Subtropical Gyre: The domain shift hypothesis. Deep-Sea Research II 48: 1449-1470.

Karl DM, Bjllrkman KM, Dore JE, Fujieki L, Hebel DV, Houlihan T, Letelier RM, Tupas LM. 2001 b. Ecological nitrogen-ro-phosphorus stoichiometry at Station ALOHA. Deep-Sea Research II 48: 1529-1566.

Karl DM, Dobbs FC. 1998. Molecular approaches to microbial biomass estimation in the sea. IN: Cooksey KE, editor. Molecular approaches to the study o/the ocean. London: Chapman & Hall. p 29-89.

Karl DM, Holm-Hansen O. 1978. Methodology and measurement of adenylate energy charge ratios in environmental samples. Marine Biology 48: 185-197.

Karl DM, Letelier RM, Hebel DV, Bird DF, Winn CD. 1992. Trichodesmium blooms and new nitrogen in the north Pacific gyre. In: Carpenter EJ et aI. (Eds), Marine Pelagic Cyanobacteria: Trichodesmium and other Diazotrophs. Kluwer Academic Publishers, Netherlands, 219-237.

Karl DM, Letelier RM, Tupas LM, Dore JE, ChristianJ, Hebel DV. 1997. The role of nitrogen fixation in biogeochemical cycling in the subtropical North Pacific Ocean. Nature 388: 533-538.

Karl DM, Lukas R. 1996. The Hawaii Ocean Time-series (HOn program: background, rationale and field implementation. Deep-Sea Research II 43: 129-156.

Karl DM, Michaels A, Bergman B, Capone D, Carpenter E, Letelier R, Lipschultz F, Paerl H, Sigman D, Stal L. 2002. Dinitrogen fixation in the world's oceans. Biogeochemistry, 57/58: 47-98.

84 Karl DM, Tien G. 1992. MAGIC: a sensitive and precise method for measuring dissolved phosphorus in aquatic environments. Limnology and Oceanography 37: 105-116.

Karl DM, Tien G. 1997. Temporal variability in dissolved phosphorus concentrations in the subtropical North Pacific Ocean. Marine Chemistry 56: 77-96.

Kiro'boe T, Hansen JLS. 1993. Phytoplankton aggregate formation - observations of patterns and mechanisms of cell sticking and the significance of exopolymeric material. Journal of Plankton Research 15: 993-1018.

Landry MR, and Kirchman D1. 2002. Microbial community structure and variability in the tropical Pacific. Deep-Sea Research II 49: 2669-2693.

Landry MR, Kirshtein J, Constantinou J. 1996. Abundances and distributions of picoplankton populations in the central equatorial Pacific from 12' N to 12' S, 140' W. Deep-Sea Research II 43: 871-890.

Landry MR et al. 1997. Iron and grazing constraints on primary production in the central equatorial Pacific: An EqPac synthesis. Limnology and Oceanography 42: 405-418.

Landry MR, Constantinou J, Latasa M, Brown SL, Bidigare RR, Ondrusek ME. 2000a Biological response to iron fertilization in the eastern equatorial Pacific (IronEx II). m., Dynamics of phytoplankton growth and microzooplankton grazing. Marine Ecology Progress Series 201: 57-72.

Landry MR, Ondrusek ME, Tanner SJ, Brown SL, Constantinou J, Bidigare RR, Coale KH, Fitzwater S. 2000b. Biological response to iron fertilization in the eastern equatorial Pacific (IronEx II) I., Microplankton community abundances and biomass. Marine Ecology Progress Series 201: 27-42.

Landry MR, Brown SL, Neveux J, Dupouy C, Blanchot J, Christensen S, Bidigare RR. 2003. Phytoplankton growth and microzooplankton grazing in high-nutrient, low­ chlorophyll waters of the equatorial Pacific: Community and taxon-specific rate assessments from pigment and flow cytometric analyses. Journal of Geophysical Research 108(CI2), 8142, doi: 10.1029/2000JCOO0744.

Latasa M, Landry MR, Schluter L, Bidigare RR. 1997. Pigment-specific growth and grazing rates of phytoplankton in the central equatorial Pacific. Limnology and Oceanography 42: 289-298.

Laws EA, Bannister IT. 1980. Nutrient-and light-limited growth of Thalassiosira fluviatilis in continuous culture, with implications for phytoplankton growth in the ocean. Limnology and Oceanography 25:457-473.

85 Le Borgne R, Barber RT, Delcroix T, Inoue HY, Mackey OJ, Rodier M. 2002. Pacific warm pool and divergence: temporal and zonal variations on the equator and their effects on the biological pump. Deep-Sea Research 1149: 2471-2512.

Le Bouteiller A. Blanchot J, Rodier M. 1992. Size distribution patterns of phytoplankton in the western Pacific: Towards a generalization for the tropical open ocean. Deep-Sea Research 39: 805-823.

Legendre L, Le Fevre J. 1989. Hydrodynamical singularities as controls ofrecyc1ed versus export production in oceans. In Berger WH, Smetacek VS, Wefer G, editors. Productivity ofthe ocean: present and past. New York: John Wiley & Sons. p 49-63.

Letelier RM, Karl DM. 1996. Role of Trichodesmium spp. in the productivity of the subtropical North Pacific Ocean. Marine Ecology Progress Series 133: 263-273.

Li WKW, SubbaRao V, Harrison WG, SmithJC, Cullen JJ, Irwin B, Platt T. 1983. Autotrophic picoplankton in the tropical ocean. Science 219: 292-295

Li WKW, Harrison WG. 2001. Chlorophyll, bacteria and picoplankton in ecological province of the North Atlantic. Deep-Sea Research II 48: 2271-2293.

Liebig J. 1840. Chemistry in its application to agriculture and physiology. Taylor and Walton

Longhurst AR. 1993. Seasonal cooling and blooming in tropical oceans. Deep-Sea Research 40: 2145-2165.

Longhurst AR 1998. Ecological Geography ofthe Sea, Academic Press, London.

Luther DS, Johnson ES. 1990. Eddy energetics in the upper equatorial Pacific during the Hawaii-to-Tahiti Shuttle Experiment. Journal of Physical Oceanography 20: 913- 944.

Mackey OJ, Blanchot J, Higgins HW, Neveux J. 2002. Phytoplankton abundances and community structure in the equatorial Pacific. Deep-Sea Research 1149:2561- 2582.

Mackey OJ, Higgins HW, Mackey MD, Holdsworth D. 1998. Algal class abundances in the western equatorial Pacific: estimation form HPLC measurements of pigments using CHEMTAX. Deep-Sea Research I 45: 1441-1468.

Marafi6n E, Holligan PM, Varela M, Mourii'l.o B, Bale AJ. 2000. Basin-scale variability of phytoplankton biomass, production and growth in the Atlantic Ocean. Deep­ Sea Research I 47: 825-857.

86 Marafl6n E, Holligan PM, Barciela R. Gonzalez N, Mouriflo B, Paz6 MJ, Varela M. 2001. Patterns of phytoplankton size structure and productivity in contrasting open ocean environments. Marine Ecology Progress Series 216: 43-56.

Marafl6n E, Behrenfeld MJ, Gonzalez N, Mouriflo B, Zubkov MV. 2003. High variability of primary production in oligotrophic waters of the Atlantic Ocean: uncoupling from phytoplankton biomass and size structure. Marine Ecology Progress Series 257: 1-11.

MargalefR. 1978. Life-fonns of phytoplankton as survival alternatives in an unstable environment. OceanologicaActa 1: 493-509.

Martin JR. 1990. Glacial-interglacial C(h change: The iron Hypothesis. Paleoceanography 5: 1-13.

Martin JH, Fitzwater SE. 1988. Iron deficiency limits phytoplankton growth in the north-eas1Pacific subarctic. Nature 331: 341-343.

McCarthy JJ, Garside C, Nevins JL, Barber RT. 1996. New production alontgl40· Win the equatorial Pacific during and following the 1992 El Nino event Deep-Sea Research II 43: 1065-1093.

Michaels AF, Silver MW. 1988. Primary production. sinking fluxes and the microbial food web. Deep Sea Research 35: 473-490.

Minas HJ, Minas M, Packard IT. 1986. Productivity in upwelling areas deduced from hydrographic and chemical fields. Limnology and Oceanography 31: 1182-1206.

Monger BC, Landry MR. 1993. Flow cytometric analysis of marine bacteria with Hoechst 33342. Applied and Environmental Microbiology. 59: 905-911.

Moore LR. Goericke R. Chisholm SW. 1995. Comparative physiology of Synechococcus and Prochlorococcus: influence of light and temperature on growth, pigments, fluorescence and absorptive properties. Marine Ecology Progress Series 116: 259-275.

Murphy LS, Haugen EM. 1985. The distribution and abundances of phototrophic ultraplankton in the North Atlantic. Limnology and Oceanography 30: 47-58.

Murray JW, Downs IN, Strom S, Wei C-L, Jannasch HW. 1989. Nutrient assimilation, export production and ~ scavenging in the eastern equatorial Pacific. Deep Sea Research 36: 1471-1489.

87 Neveux J, Lantoine F, Vaulot D, Marie D, BlanchotJ. 1999. Phycoerytbrins in the southern tropical and equatorial Pacific Ocean: Evidence for new cyanobacterial types. Journal of Geophysical Research 104: 3311-3322.

Neveux J, Dupouy C, Blanchot J, Ie Bouteiller A, Landry MR, Brown SL. 2003. Diel dynamics of chlorophylls in high-nutrient, low-chlorophyll waters of the equatorial Pacific (l80'): Interactions of growth, grazing, physiological responses and mixing. Journal of Geophysical Research 108(CI2), 8140, doi: 10.1029/2000JC000747.

Panensky F, Hoepffner N, Li WKW, 0110a 0, Vaulot D. 1993. Photoacclimation of Prochlorococcus spp. (Prochlorophyta) strains isolated form the North Atlantic and the Mediterranean Sea. Plant Physiology 101: 285-296.

Partensky F, Hess WR. Vaulot D. 1999. Prochlorococcus a marine photosynthetic prokaryote of global significance. Microbiology and Molecular Biology Reviews 63: 106-107.

Pefia MA, Lewis MR, Harrison WG. 1990. Primary productivity and size structure of phytoplankton biomass on a transect of the equator at 135' W in the Pacific Ocean. Deep-Sea Research 37: 295-315.

Pefia MA, Harrison WG, Lewis MR. 1992. New production in the central equatorial Pacific. Marine Ecology Progress Series 80: 265-274.

Perez v, FernAndez E, MaraIl6n E, Serret P, Varela R, Bode A, Varela M, Varela MM, Moran ZAG, Woodward EMS, Kitidis V, Garcfa-Soto C. 2005. Latitudinal distribution of microbial plankton abundance, production, and respiration in the Equatorial Atlantic in autumn 2000. Deep-Sea Research I 52: 861-880.

Philander SG. 1990. El Nino. La Nino and the Southern Oscillation. Academic Press, San Diego.

PicautJ, Ioualalen M, Menkes C, Delcroix T, McPhaden MJ. 1996. Mechanism of the zonal displacements of the Pacific warm pool: implications for ENSO. Science 274: 1486-1489.

Platt T, Denman KL. 1 977. Organization in the pelagic ecosystem. Helgolaender wiss.Meeresunters 30: 572-581.

Poulton AJ, Holligan PM, Hickman A, Kim YN, Adey TR, Stinchcombe MC, Holeton C, Root S, Woodward EMS. 2006. Phytoplankton carbon fixation, chlorophyll­ biomass and diagnostic pigments in the Atlantic Ocean. Deep-Sea Research II 53: 1593-1610.

88 Price NM, Anderson LF, Morel FMM. 1991. Iron and nitrogen nutrition of equatorial Pacific plankton. Deep Sea Research 38: 1361-1378.

Raven JA. 1986. Physiological consequences of extremely small size for autotrophic organisms in the sea. In Platt T, Li WKW, editors, Photosynthetic picoplankton, Can Bull Fish Aquat Sci 214: 1-70.

Redfield AC, Ketchum BH, Richards FA. 1963. The influence of organisms on the composition of seawater. In: Hill MN, editor. The sea: ideas and observations on progress in the study ofthe seas. New York: Interscience. p 26-77.

Rhee G-Y. 1978. Effects ofN:P atomic ratios and nitrate limitation on algal growth, cell composition and nitrate uptake. Limnology and Oceanography 23:10-25.

Richardson TL, Jackson GA. 2007. SmaIl phytoplankton and carbon export from the surface ocean. Science 315: 838-840.

Rippka R, Coursin T, Hess W, Lichtle C, Scanlan DJ, Palinska KA, Iternan I, Partensky F, Houmard J, Herdman M. 2000. Prochlorococcus marinus Chisholm et aI., 1992, subsp. Nov. pastoris, strain PCC 9511, the first axenic chlorophyll a2ibr containing cyanobacterium (Oxyphotobacteria). International Journa1 of Systematic and Evolutionary Microbiology 50: 1833-1847.

Roden GI. 1975. On North Pacific temperature, salinity, sound velocity and density fronts and their relation wind and energy flux fields. Journa1 of Physical Oceanography 5: 557-571.

Rodier M, Eldin G, Le Borgne R. 2000. The western boundary of the equatorial Pacific upwelling: some consequences of climatic Variability on hydrological and planktonic properties. Journa1 of Oceanography 56: 463-471.

Sakshaug E, Holm-Hansen O. 1977. Chemical composition of Skeletonema costalum (Grev.) Cleve and Pavlova (Monochrysis) lutheri (Droop) Green as a function of nitrate-, phosphate-, and iron-limited growth. Journa1 of Experimental Marine Biology and Ecology 29: 1-34.

Salihoglu B, Hofinann EE. 2007. Simulations of phytoplankton species and carbon production in the equatorial Pacific Ocean 2. Effects of physical and biogeochemical processes. Journa1 of Marine Research 65: 275-300.

San Martin E, Harris RP, Irigoien X. 2006. Latitudinal variation in plankton size spectra in the Atlantic Ocean. Deep-Sea Research II 53: 1560-1572.

Sheldon RW, Prakash A, Sutcliffe Jr. WH. 1972. The size distribution of particles in the ocean. Limnology and Oceanography 17: 327-340.

89 Sheldon RW, Sutcliffe Jr. WH, Paranjae MA. 1977. Structure of pelagic food chain and relationship between plankton and fish production. Journal of the Fisheries Research Board Cananda 34: 2344-2353.

Sieburth J MeN, Smetacek V, Lenz J, 1978. Pelagic ecosystem structure: heterotrophic compartments of plankton and their relationship to plankton size fractions. Limnology and Oceanography 23:1256-63.

Smith CR, Hoover OJ, Doan SE, Pope RH, Demaster OJ, Dobbs FC, Altabet MA. 1996. Phytodetritus at the abyssal seafloor across 10' of latitude in the central equatorial Pacific. Deep Sea Research 43: 1309-1338.

Stewart DE, Farmer FH. 1984. Extraction, identification and quantification of phycobiliprotein pigments from phototrophic plankton. Limnology and Oceanography 29: 392-397.

Stoens A, Menkes C, Radenac MH, Dandonneau Y, Grima N, Eldin G, Memery L, Navarette C, Andre JM, Moutin T, Raimbault P. 1999. The coupled physical­ new production system in the equatorial Pacific during the 1992-1995 EI Niiio. Journal of Geophysical Research 104: 3323-3339.

Strickland JDH, Parsons TR. 1972. A Practical handbook a/Seawater Analysis, Fisheries Research Board of Canada, Ottawa.

Takahashi M, Bienfang PK. 1983. Size structure of phytoplankton biomass and photosynthesis in subtropical Hawaiian waters. Marine Biology 76: 203-211.

Tans PP, Fung IY, Takahashi T. 1990. Observational constraints on the global atmospheric carbon dioxide budget Science 247: 1431-1438.

Tett P, Heaney SI, Droop MR. 1985. The Redfield ratio and phytoplankton growth rete. J Mar BioI Assoc Uk 65:487-504.

Thomas WH. 1972. Nutrient inversions in the southeastern tropical Pacific Ocean. Fisheries Bulletin 70: 929-933.

Thomas WH. 1979. Anomalous nutrient-chlorophyll interrelationships in the offshore eastern tropical Pacific Ocean. Journal of Marine Research 37: 327-335.

Thomas-Bulldis, Karl DM. 1998. Application of a novel method for phosphate and non­ phosphate phosphorus determinations in the oligotrophic North Pacific Ocean. Limnology and Oceanography 43: 1565-1577.

90 Tremblay JE, Legendre L. 1994. A model for the size-fractionated biomass and production of marine phytoplankton. Limnology and Oceanography 39: 2004- 2014.

Vidondo B, Prairie YT, Blanco JM, Duarte CM. 1997. Some aspects of the analysis of size spectra in aquatic ecology. Limnology and Oceanography 42: 184-192.

Vidussi F, Claus1re H, Manca BB, Luchetta A, Marty JC. 2001. Phytoplankton pigment distribution in relation to upper thermocline circulation in the eastern Mediterranean Sea during winter. Journal of Geophysical Research 106: 19939- 19956.

VIlIrS N, Buck KR, Chavez FP, Eikrem W, Hansen LE, Ostergaard JB, Thomsen HA. 1995. Nanoplankton of the equatorial Pacific with emphasis on heterotrophic protists. Deep-Sea Research 42: 585-602.

Walsh JJ. 1976. Herbivory as a factor in patterns of nutrient utilization in the sea. Limnology and Oceanography 21: 1-13.

Waterbury JB, Watson SW, Gullard RRL, Brand LE. 1979. Widespread occurrence ofa unicellular cyanobacterium Synechococcus. Canadian Journal of Fisheries and Aquatic Science 214: 71-120.

Waterbury JB, Watson SW, Balois FW, Franks DG. 1986. Biological and ecological characterization of the marine unicellular cyanobacteria Synechococcus. Canadian Bulletin ofFisheries and Aquatic Sciences 214: 71-120.

Wilkerson FP, Dugdal RC. 1996. Silicate versus nitrate limitation in the equatorial Pacific estimated from satellite-derived sea-surface temperatures. Advances in Space Research 18: 81-89.

Williams PJ LeB. 1981. Incorporation of micro heterotrophic processes in the classical paradigm of the planktonic food web. Kieler Meeresforschungen. Sonderheft 5: 1- 28.

Winn CD, Campbell L, Christian JR, Letelier RM, Hebel DV, Dore JE, Fujieki L, Karl DM. 1995. Seasonal variability in the phytoplankton community of the North Pacific Subtropical Gyre. Global Biogeochemical Cycles 9: 605-620.

Wright SW, Jeffrey SW. Mantoura RFC, Lewellyn CA, Bjornland T, Repeta D, Welschmeyer N. 1991. Improved HPLC method for the analysis of chlorophylls and carotenoids from marine phytoplankton. Marine Ecology Progress Series 77: 183-96.

91 Wyman M. 1992. An in vivo method for the estimation of phycoerythrin concentrations in marine cyanobacteria (Synechococcus spp.). Limnology and Oceanography 37:1300-1306.

Zehr JP, Mellon MT, Zani S. 1998. New nitrogen-fixing microorganisms detected in oligotrophic oceans by amplification of nitrogenase (nifH) genes. Applied and Environmental Microbiology 64: 3444-3450.

Zher JP, Waterbury JB, Turner PJ, Montoya JP, Omoregie E, Steward OF, Hansen A, Karl DM. 2001. Unicellular cyanobacteria fix N, in the subtropical North Pacific Ocean. Nature 412: 635-638.

Zubkov MV, Sleigh MA, Tarran OA, Burkill PH, Leakey RJO. 1998. Picoplanktonic community structure on an Atlantic transect from 50· N to 50· S. Deep-Sea Research I 45: 1339-1355.

Zubkov MV, Sleigh MA, Burkill PH. 2000. Assaying picoplankton distribution by flow cytometry of underway samples collected along a meridional transect across the Atlantic Ocean. Aquatic Microbial Ecology 21: 13-20.

92