• No results found

Late Quaternary climatic and oceanographic changes in the Northeast Pacific as recorded by dinoflagellate cysts from Guaymas Basin, Gulf of California (Mexico)

N/A
N/A
Protected

Academic year: 2021

Share "Late Quaternary climatic and oceanographic changes in the Northeast Pacific as recorded by dinoflagellate cysts from Guaymas Basin, Gulf of California (Mexico)"

Copied!
108
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

by dinoflagellate cysts from Guaymas Basin, Gulf of California (Mexico)

by

Andrea Michelle Price

B.Sc., University of Victoria, 2010

A Thesis Submitted in Partial Fulfillment

of the Requirements for the Degree of

MASTER OF SCIENCE

in the School of Earth and Ocean Sciences

 Andrea Michelle Price, 2012

University of Victoria

All rights reserved. This thesis may not be reproduced in whole or in part, by photocopy

or other means, without the permission of the author.

(2)

ii

Supervisory Committee

Late Quaternary climatic and oceanographic changes in the Northeast Pacific as recorded

by dinoflagellate cysts from Guaymas Basin, Gulf of California (Mexico)

by

Andrea Michelle Price

B.Sc., University of Victoria, 2010

Supervisory Committee

Dr. Vera Pospelova, (School of Earth and Ocean Sciences)

Supervisor

Dr. Thomas Pedersen, (School of Earth and Ocean Sciences)

Departmental Member

Dr. Richard Hebda, (School of Earth and Ocean Sciences, Department of Biology)

(3)

iii

Abstract

Supervisory Committee

Dr. Vera Pospelova (School of Earth and Ocean Sciences)

Supervisor

Dr. Thomas Pedersen (School of Earth and Ocean Sciences)

Departmental Member

Dr. Richard Hebda (School of Earth and Ocean Sciences, Department of Biology)

Departmental Member

A high-resolution record of organic-walled dinoflagellate cyst production in

Guaymas Basin, Gulf of California (Mexico) reveals a complex paleoceanographic

history over the last ~40 ka. Guaymas Basin is an excellent location to perform high

resolution studies of changes in Late Quaternary climate and paleo-productivity because

it is characterized by high primary productivity, high sedimentation rates, and low

oxygen bottom waters. These factors contribute to the deposition and preservation of

laminated sediments throughout large portions of the core MD02-2515. In this study we

document dinoflagellate cyst production at a centennial to millennial scale throughout the

Late Quaternary. Based on the cyst assemblages three major dinoflagellate cyst zones,

with seven subzones were established. The most dominant dinoflagellate cyst taxa found

throughout the core were Brigantedinium spp. and Operculodinium centrocarpum.

Dansgaard-Oeschger events 5-8 are inferred in the dinoflagellate cyst records on the basis

of increases in warm taxa, such as Spiniferites pachydermus. Preceding and during the

Last Glacial Maximum cysts of Polykrikos cf. kofoidii increase in abundance, responding

to oceanographic changes in the Gulf of California perhaps caused by a regression in

sea-level. Other intervals of interest are the Younger Dryas where cooler conditions are not

recorded, and the Holocene which is characterized by the consistent presence of warm

water species Stelladinium reidii, Tuberculodinidum vancampoae, Bitectatodinium

spongium and an increase in Quinquecuspis concreta. Changes in cyst assemblages,

concentrations and species diversity, along with geochemical data reflect major

millennial scale climatic and oceanographic changes.

(4)

iv

Table of Contents

Supervisory Committee ... ii

Abstract... iii

Table of Contents ... iv

List of Tables ...v

List of Figures... vi

List of Plates ... viii

Acknowledgements ...x

Author Contributions ... xi

1. Introduction...1

2. Environmental Setting...5

2.1 Physiography ...5

2.2 Modern Oceanographic and Climatic Setting of the Gulf of California ...7

3. Material and Methods ...11

3.1 Palynological Sample Preparation and Microscopy ...15

3.2 Species Diversity and Multivariate Statistical Analyses ...25

4. Results ...25

4.1 Dinoflagellate Cyst Relative Abundances, Concentrations and Accumulation Rates ...25

4.2 Statistical Analyses and Dinoflagellate Cyst Zones ...29

4.3 The Modern Analogue Technique ...32

5. Discussion ...32

5.1 The (Paleo-)productivity Signal ...33

5.2

Dinoflagellate Cyst Preservation ...34

5.3 Dinoflagellate Cyst Zones 1a to 1d (~40-22.8 ka) ...34

5.3.1 Dansgaard-Oeschger events ...34

5.3.2 Preceding the Last Glacial Maximum (27.7 - 22.8 ka) ...38

5.4 Dinoflagellate Cyst Zones 2a - 2c (22.8-11.1 ka)...39

5.4.1 The Last Glacial Maximum (~22.3-17 ka)...39

5.4.2 Heinrich Event 1 (17-16 ka)...42

5.4.3 The Bølling-Allerød (14.6-12.9 ka) and Younger Dryas (12.9-11.6 ka) ...44

5.5 Dinoflagellate Cyst Zone 3a (11.1-6.8 ka) ...46

5.5.1 The Early to Mid Holocene (~11.6-6.8 ka) ...47

6. Conclusions...49

References...52

Appendix 1...67

(5)

v

List of Tables

Table 1. Taxonomic citation of dinoflagellate cysts identified in this study. Thecate

equivalents are taken from Head [1996], Zonneveld [1997], Zonneveld and Jurkschat

[1999], Marret and Kim [2009], Matsuoka et al., [2009], and Verleye et al. [2011]…... 17

(6)

vi

List of Figures

Figure 1. (A) Map of the study area showing the locations of MD02-2515 (this study)

and other nearby cores, GGC-55/JPC-56 and DSDP-480. Bathymetric contours are in

metres. Guaymas Basin (GB) is located near the center of the Gulf. (B) The east tropical

Pacific. Approximate locations of the present day Intertropical Tropical Convergence

Zone (ITCZ) (dashed line), moisture sources (solid colored arrows), North Pacific High

(NPH) (solid oval), Sonoran Desert Low (L) (dashed ovals), and prevalent wind direction

in the Gulf (doubled-headed arrow). Features shown in red correspond to northern

hemisphere (NH) summer locations or sources (for moisture), while features shown in

blue correspond to NH winter locations………. 6

Figure 2. Cross section of the Gulf of California showing water masses and their

dominant flow direction [modified after Bray, 1988a]. Dashed blue arrows show the

predominant depth at which upwelling occurs.……… 9

Figure 3. Age-depth diagram for the 17 AMS

14

C dates used to construct the age model

by Pichevin et al., [2012]. Interstadials are shown by shaded horizontal bars and

dinoflagellate cyst zones are separated by dashed lines. Heinrich events and the Younger

Dryas are highlighted in dark grey……… 13

Figure 4. GISP II δ

18

O [Stuiver and Grootes, 2000], % total organic carbon (% TOC), %

biogenic silica (% BioSi), magnetic susceptibility [Cheshire et al., 2005], sedimentation

rate, lithology and structure [Beaufort et al., 2002]. Bold curves are 10 point averages.

Interstadials are shown by shaded horizontal bars and dinoflagellate cyst zones are

separated by dashed lines. Heinrich events and the Younger Dryas are highlighted in dark

grey.……….. 14

Figure 5. Relative abundances (%) of selected dinoflagellate cyst taxa. Dinoflagellate

cysts species that characterize particular zones are highlighted in color. Interstadials are

shown by shaded horizontal bars and dinoflagellate cyst zones are separated by dashed

lines. Heinrich events and the Younger Dryas are highlighted in dark grey…………... 28

Figure 6. Sedimentation rate; concentrations and fluxes for total, autotrophic and

heterotrophic taxa; ratio of heterotrophic (H) to autotrophic (A) taxa, and A/H;

Shannon-Wiener index with colored vertical lines highlighting distinct changes; and sample scores

for PCA axes 1-4, where different colors highlight changes between zones. Interstadials

are shown by shaded horizontal bars and dinoflagellate cyst zones are separated by

dashed lines. Heinrich events and the Younger Dryas are highlighted in dark

grey………... 30

Figure 7. Ordination diagram generated by principal components analysis (PCA)

showing dinoflagellate cyst taxa (black arrows) and ordination of samples (colored dots

and triangles). Abbreviations: Bspp, Brigantedinium spp; Proto, Protoperidinium spp;

Dubr, Dubrinidium spp.; CystA, Cyst type A; SBC, spiny brown cysts and Echinidinium

(7)

vii

species; CystL, Cyst type L; Psch, cyst of Polykrikos schwartzii; Pkof, cyst of Polykrikos

kofoidii; Pcfk, cyst of Polykrikos cf. kofoidii; Ocent, Operculodinium centrocarpum

sensu Wall and Dale 1966; Sben, Spiniferites bentorii; Spac, Spiniferites pachydermus;

Nlab, Nematosphaeropsis labyrinthus; Lmac, Lingulodinium machaerophorum; Eacu,

Echinidinium aculeatum; Tvan, Tuberculodinium vancampoae; Stel, Stelladinium reidii;

Qcon, Quinquecuspis contreta; Operculodinium aguinawense; Pdal, cyst of

Pentapharsodinium dalei; Pzoh, Polysphaeridium zoharyi; Sram, includes Spiniferites

ramosus and Spiniferites bulloideus; Ssp1, Spiniferites type 1; Sspp, Spiniferites spp.;

Smir, Spiniferites mirablis; Squa, includes Selenopemphix quanta and cysts of

Protoperidinum nudum; Sneph, Selenopemphix nephroides; Sund, Selenopemphix

undulata; Imap; Impagidinium spp. Eigenvalues are shown at the bottom of the figure.

(8)

viii

List of Plates

Plate 1. Bright-field photomicrographs. Scale bars are 20 µm. 1. Impagidinium

aculeatum, UVic 11-060, slide 1. 2.-3. Impagidinium paradoxum, UVic 06- 578, slide 1.

4.-5. Impagidinium sphaericum, UVic 06-597, slide 1. 6, 9. Impagidinium strialatum,

UVic 06-612, slide 1. 7-8. Nematosphaeropsis labyrinthus, UVic 06-613, slide 1……. 18

Plate 2

Bright-field photomicrographs. Scale bars are 20 µm. 1. Lingulodinium

machaerophorum, arrow shows granules on the tip of the process, UVic 11-039, slide 1.

2. Cyst of Pentapharsodinium dalei, UVic 06-576, slide 2. 3. Polysphaeridium zoharyi

showing epicystal archeopyle, UVic 11-143, slide 1. 4-5. Operculodinium centrocarpum

sensu Wall and Dale 1966 showing precingular archeopyle, UVic 06-610 slide 2, UVic

06-621, slide 1. 6. Operculodinium centrocarpum var. truncatum, UVic 06-610, slide 2.

7-8. Operculodinium aguinawense, arrow shows striated base of processes, UVic 06-634,

slide 1. 9. Operculodinium type 1., UVic 06-639, slide 1……… 19

Plate 3. Bright-field photomicrographs. Scale bars are 20 µm. 1. Spiniferites belerius,

UVic 06-639, slide 1. 2. Spiniferites bentorii, UVic 06-568, slide 1. 3.-4. Spiniferites

bentorii with reduced processes, UVic 610, slide 2. 5-6. Spiniferites type 1, UVic

06-576, slide 1.7. Spiniferites bulloideus, UVic 06-566, slide 2. 8. Spiniferites ramosus,

UVic 11-132, slide 1. 9. Spiniferites pachydermus, UVic 11-039, slide 1. 10. Spiniferites

mirabilis, UVic 11-040, slide 1. 11.-12. Spiniferites cf. hyperacanthus, UVic 06-610-2,

slide 2……… 20

Plate 4. Bright-field photomicrographs. Scale bars are 20 µm. 1. Bitectatodinum

spongium, UVic 06-619, slide 1. 2. Tectadinium pellitum, UVic 06-581, slide 1. 3.

Tuberculodinium vancampoae, UVic 11-058, slide 1. 4. Brigantedinium simplex, UVic

11-045, slide 1. 5. Brigantedinium cariacoense, UVic 11-058, slide 1. 6. Brigantedinium

spp., UVic 06-665, slide 1. 7. Dubridinium spp. UVic 06-560, slide 1. 8. Cyst of

Protoperidinium americanum, UVic 06-649, slide 1. 9. Gymnodinium spp., polygon in

white highlights the reticulation, UVic 11-134, slide 1……… 22

Plate 5. Bright-field photomicrographs. Scale bars are 20 µm. 1. Quinquecuspis

concreta, UVic 06-671, slide 1. 2. Cyst of Protoperidinium oblongum sensu Wall and

Dale 1968, UVic 11-041, slide 1. 3. Votadinium calvum, UVic 06-665, slide 1. 4.

Votadinium spinosum, UVic 06-558, slide 1. 5. Cyst of Protoperidinium nudum, UVic

11-059, slide 1. 6. Selenopemphix quanta, UVic 06-650, slide 1. 7. Selenopemphix

nephroides, UVic 11-041, slide 1. 8. Selenopemphix undulata, UVic 06-566, slide 1. 9.

Stelladinium reidii, UVic 11-057, slide 1………. 23

Plate 6. Bright-field photomicrographs. Scale bars are 20 µm. 1. Cyst of Polykrikos

kofoidii UVic 06-670, slide 1. 2. Cyst of Polykrikos schwartzii, UVic 06-611, slide 1.

3. Cyst of Polykrikos cf. kofoidii, UVic 06-552, slide 2. 4. Echinidinium aculeatum, UVic

06-543, slide 1. 5. Echinidinium granulatum, UVic 143-1, slide 1. 6. Echinidinium cf.

(9)

ix

zonneveldiae, UVic 06-644, slide 1. 7. Cyst type A, UVic 11-042, slide 1. 8. Cyst type F,

(10)

x

Acknowledgements

First and foremost I would like to thank my supervisor and mentor Dr. Vera

Pospelova for her tremendous support, encouragement, and guidance, and for first

sparking my interest in micropaleontology. I would also like to thank Dr. Kenneth

Mertens for his contributions and feedback, my committee members Dr. Tom Pedersen

and Dr. Richard Hebda for their time and willingness to serve on my thesis committee,

and Dr. Rolf Mathewes for serving as the external examiner. I am grateful to Victoria

Gray, Alanna Krepakevich, and Kristen Kennedy for their laboratory assistance and

would like to thank Manuel Bringué for his helpfulness, interesting discussions, and

introducing me to new music genres while counting at the microscope. Thank you to

Ruth Dwyer (University of Edinburgh) for sampling the core, Dr. Heather Cheshire

(University College London) for providing sediment density and magnetic susceptibility

measurements, Ellen Roosen (Woods Hole Oceanographic Institute) for providing

surface sediment samples and Dr. Terri Lacourse (University of Victoria) for discussions

about multivariate statistics. Finally, I would like to thank my family and friends for their

unwavering love and support. Funding for this project was provided by the Natural

Sciences and Engineering Council of Canada (NSERC) through grants to Dr. Vera

Pospelova, Dr. Tom Pedersen and a CGS-M scholarship to Andrea Price; a UVic

Graduate Fellowship, a UVic David Strong Research Scholarship, a UVic President’s

Research Scholarship, a Martlet Chapter IODE Graduate Scholarship for Women, and an

American Association of Stratigraphic Palynologists (AASP) Student Scholarship to

Andrea Price.

(11)

xi

Author contributions

This thesis forms the basis of a manuscript that has been submitted (June 2012) to

the journal Paleoceanography coauthored by Andrea M. Price, Kenneth N. Mertens,

Vera Pospelova, Thomas F. Pedersen and Raja S. Ganeshram.

Andrea M. Price – processed 47 samples, microscopic analysis of 178 samples,

preformed statistical analyses, made maps and figures, analyzed and interpreted the data,

and prepared the manuscript for publication.

Kenneth N. Mertens – microscopic analysis of 108 additional samples and provided

detailed comments, suggestions and corrections on the manuscript.

Vera Pospelova – initiation of the project and acquisition of samples, funding through

NSERC grants, provided advice and input into the interpretation of the data as well as

comments, suggestions and corrections on the manuscript.

Thomas F. Pedersen – provided partial funding for laboratory processing of sediments.

Raja S. Ganeshram – provided samples.

(12)

1. Introduction

Abrupt millennial scale climate fluctuations have occurred during the last glacial,

namely the quasiperiodic Dansgaard-Oeschger (D-O) events. They are best documented

by δ

18

O records from the Greenland ice sheet [i.e. Dansgaard et al., 1993], where

positive excursions mark warm interstadials and occur on average ~1500 years apart

[Clement and Peterson, 2008]. During some extreme cold stadials ice rafted debris was

deposited in the North Atlantic, and these have been termed Heinrich events [Broecker et

al., 1992]. Millennial scale variability is most pronounced in the North Atlantic, but is

also observed in paleoclimate records from lower latitudes including the tropics [Voelker

et al., 2002 and references within; Clement and Peterson, 2008]. The apparent

synchronicity and global nature of these events suggests a tight coupling between shifts

in atmospheric and ocean circulation over broad areas [Behl and Kennett, 1996; Hendy

and Kennett, 1999]. Regions in the northern tropics are particularly sensitive due to

latitudinal shifts in the Intertropical Convergence Zone (ITCZ) and related shifts in

precipitation patterns [e.g. Peterson et al., 2000]. The latitudinal position of the ITCZ is

influenced by seasonal differences in temperature and incoming solar insolation, as well

as El Niño-Southern Oscillation (ENSO) variability on interannual timescales. During

warm episodes the pole-to-Equator gradient is lessened, shifting the ITCZ further north,

while the opposite occurs during cold periods [e.g. Asmerom et al., 2010].

The Gulf of California (herein referred to as the Gulf) is an excellent location to

investigate millennial scale variability during the last glacial period because it has a

monsoonal climate that is heavily modulated by the location of the ITCZ, and nearby

pressure systems such as the North Pacific High (NPH) and Sonoran Desert Low (SDL)

(13)

2

(Figure 1b). Guaymas Basin, located in the central Gulf, has the potential to be an ideal

recorder of paleoceanographic information because it is characterized by low oxygen

bottom water, seasonal patterns in upwelling and precipitation, and high primary

productivity which contribute to the deposition and preservation of laminated sediments.

Guaymas Basin has been the focus of Late Quaternary paleoclimate reconstructions,

especially over the past ~17 ka [e.g. Keigwin and Jones, 1990; Sancetta, 1995; Pride et

al., 1999; Barron et al., 2004; Dean, 2006]. These studies have provided valuable

information on past climatic and oceanographic conditions in the Northeast Pacific since

the Last Glacial Maximum (LGM). For example, Pride et al. [1999] studied nitrogen

isotopes and biogenic silica (BioSi) records from cores JPC56 and GCC55 in Guaymas

Basin and found laminated sediments with high δ

15

N

org

and BioSi accumulation rates

during the Bølling-Allerød (B-A), while the Younger Dryas (YD) and end of the late

glacial were characterized by non-laminated sediments, low BioSi, and lower δ

15

N

org

.

They attributed these shifts to widespread changes in the extent of suboxic subsurface

waters, variations in upwelling and mixing, and/or the latitudinal migration of the ITCZ.

More recently Barron et al. [2004] analyzed geochemical, diatom and silicoflagellate

records from DSDP Site 480 from ~15 - 1 ka. Like Pride et al. [1999] they found that the

B-A was characterized by diatom rich laminated sediments, while the YD saw a decline

in biogenic silica. Barron et al. [2004] also documented an increase in CaCO

3

, and the

presence of tropical diatom species that suggest reduced upwelling conditions during the

YD. In addition, there are few well-dated cores in the Gulf that extend past ~17 ka.

In this study we use dinoflagellate cysts to investigate millennial scale variability

in primary productivity and climate over the past 40 ka BP. Dinoflagellates are one of the

(14)

3

most important groups of primary producers in coastal and estuarine systems, with over

1,555 known free-living marine species worldwide [Gómez, 2005]. They are

predominately single-celled protists, and are characterized by two flagella, the presence

of unique sterols, lack of c

1

chlorophyll, a unique cell wall structure and characteristic

nuclei [Dale, 1996; Fensome et al., 1996; Mudie and Harland, 1996]. Furthermore they

display a high degree of diversity both in their motile stage [Sarjeant et al., 1987] and

resting stage.

They are unique amongst other phytoplankton groups in several respects. Firstly,

dinoflagellates display a high degree of mobility. Although coccolithophores also have a

motile stage, and diatoms to some extent are able to adjust their position in the water

column, dinoflagellates actively swim and are thus more effectively mobile [Dale, 1996].

This gives them many advantages as they are able to maximize photosynthesis by

regulating their depth in the water column [Prezelin, 1987] and they can maintain their

depth at nutrient-rich oceanic fronts [Dale, 1996]. They are especially adapted to well

stratified waters where they are able to photosynthesize near the surface during the day

and migrate down into higher nutrient subsurface waters at night, thus making swimming

biologically worthwhile. Increased stratification is less favorable to diatoms and the

heterotrophic dinoflagellates that feed on them, resulting in less competition for nutrients

and may result in an increase in autotrophic taxa. Heterotrophic dinoflagellates on the

other hand tend to dominate in upwelling conditions when diatoms are abundant [i.e.

Taylor, 1987]. Secondly, dinoflagellates employ a variety of trophic strategies.

Approximately half of all dinoflagellates are heterotrophs, while many others may use

more than one nutritional strategy (mixotrophy), rather than being obligatory autotrophs

(15)

4

[Dale, 2009]. The differences in trophic structure impacts species distribution. Those

that primarily rely on photosynthesis to synthesize food are most affected by nutrient

concentrations and light levels, while heterotrophs respond more to prey availability

[Dale, 1996]. Thirdly, numerous dinoflagellate species produce an organic-walled cyst, a

dormant stage, that is highly resistant to physical, chemical and biological degradation

[Dale, 1996; Fensome et al., 1996], unlike siliceous or calcareous microfossils which are

subject to dissolution. Resting cysts are formed after sexual reproduction. The formation

of cysts has a variety of potential ecological functions including survival through adverse

environmental conditions, seeding source for the motile population, bloom initiation,

species dispersal, and increased genetic recombination [Sarjeant et al., 1987].

Dinoflagellate species have different environmental preferences and over the last

few decades cyst assemblages have been developed as tools used in paleoenvironmental

reconstructions. They are used to determine past surface temperature (SST),

sea-surface salinity (SSS), primary productivity, eutrophication, pollution, and sea-ice

coverage [e.g. Dale, 1996; Matsuoka, 1999; Rochon et al., 1999; de Vernal et al., 2001;

Marret and Zonneveld, 2003; Radi and de Vernal, 2004; Radi et al., 2007; Pospelova et

al., 2008; Dale, 2009; Zonneveld et al., 2009].

This is the first study to analyze dinoflagellate cyst records at the centennial to

millennial scale in the Northeast Pacific over the Late Quaternary. We document

variations in dinoflagellate cyst species composition and abundance over the past 40 ka in

Guaymas Basin. Our objectives are (1) to describe changes in marine primary

productivity in relation to past climatic variability, (2) to determine if dinoflagellate cyst

assemblages record millennial scale climatic and oceanographic changes in the Gulf of

(16)

5

California, (3) to compare our records to other proxies of sea-surface conditions in the

Gulf and (4) to provide further insights into the oceanographic and climatic history of the

region over the Late Quaternary.

2. Environmental Setting

2.1 Physiography

The Gulf of California is a narrow marginal sea that opens into the subtropical

Pacific Ocean. It is located between the Baja Peninsula and the northwest coast of

Mexico and is approximately 1,000 km in length and 150 km in width, reaching water

depths of over 2,500 m [Bray, 1988a] (Figure 1a). Guaymas Basin is situated

approximately in the center of the Gulf and is an elongate semi-enclosed basin that

formed via seafloor spreading [Dean, 2006]. The Rio Yaqui and Rio Matape drain into

the eastern margin of Guaymas Basin, while the Colorado River, Rio Sonora, Rio Mayo

and Rio Fuertre also drain into the Gulf, but are located further north or south of

Guaymas Basin (Figure 1a). These rivers contribute some terrigenous material into the

basin, however fluvial input of terrigenous material is thought to be minor [Baba et al.,

1991; Baumgartner et al., 1991; Thunell et al., 1993].

(17)

6

Figure 1. (A) Map of the study area showing the locations of MD02-2515 (this study)

and other nearby cores, GGC-55/JPC-56 and DSDP-480. Bathymetric contours are in

metres. Guaymas Basin (GB) is located near the center of the Gulf. (B) The east tropical

Pacific. Approximate locations of the present day Intertropical Tropical Convergence

Zone (ITCZ) (dashed line), moisture sources (solid colored arrows), North Pacific High

(NPH) (solid oval), Sonoran Desert Low (L) (dashed ovals), and prevalent wind direction

in the Gulf (doubled-headed arrow). Features shown in red correspond to northern

hemisphere (NH) summer locations or sources (for moisture), while features shown in

blue correspond to NH winter locations.

Baja C alifornia

MEXICO

Colorado River Rio Yacqui Rio Matape

Pacific

Ocean

Rio Mayo Rio Fuerie Rio Sonora Guaymas 55/56480

Gulf of

California

2000 m 114 W 112 W 110 W 108 W 24 N 26 N 30 N 28 N 32 N 500 m 200 m GB 2515 NH winter ITCZ NH summer ITCZ NPH L L 0 30 N 15 N 15 S 120 W 110 W 100 W 90 W 80 W 70 W

A.

B.

Midriff Islands 100 km

(18)

7

2.2 Modern Oceanographic and Climatic Setting of the Gulf of California

The Gulf of California is the only evaporative basin adjacent to the Pacific Ocean.

It is unique in that unlike many mid-latitude evaporative basins, the Gulf of California

has net outflow at the surface (above 250 m) and inflow at depth (500-250 m) [Bray,

1988b]. Surface circulation and upwelling is strongly controlled by atmospheric systems,

where wind direction and strength determine the location and intensity of upwelling in

the Gulf [Thunell et al., 1993; Thunell et al., 1994; Douglas et al., 2007].

Water masses

There are a number of water masses present in the Gulf that impact primary

productivity and lamination preservation. The bottom water mass, the Pacific Deep Water

(PDW), is found below depths of ~1000 m in the Gulf (Figure 2). At depths between

~500-1000 m Pacific Intermediate Water (PIW) flows into the Gulf creating a persistent

Oxygen Minimum Zone (OMZ). The OMZ causes slope sediments at these depths to be

sufficiently depleted in oxygen that bioturbation is inhibited, allowing for the

preservation of laminated sediment [Cheshire et al., 2005]. The upper ~500 m involved

in thermohaline circulation consists of three layers, with inflow at 500-250 m, outflow at

250-50 m, and a surface layer (~50-0 m) that reverses seasonally with the wind direction

(Figure 2). During summer warm surface water is transported into the Gulf with

southeasterly winds (Figure 1b). In winter the opposite occurs and colder surface water is

transported out of the Gulf. Water mass formation occurs in the northern Gulf where

Subsurface Subequatorial Water (SSW) and water from the Colorado delta mix by winter

(19)

8

convection, tidal mixing, and buoyancy driven horizontal circulation, forming Northern

Gulf Water (NGW) (Figure 2) [Bray, 1988a; Pride et al., 1999]. Water mass formation in

the northern Gulf contributes significantly to high productivity in Guaymas Basin

[Baumgartner, 1987]. Upwelled water originates from the Central Gulf Water (CGW), at

a water depth of 250-50 m; thus the nutrient content of this water mass can have a

(20)

9

Figure 2. Cross section of the Gulf of California showing water masses and their

dominant flow direction [modified after Bray, 1988a]. Dashed blue arrows show the

predominant depth at which upwelling occurs.

Guaymas

Basin

NW

SE

PDW

PIW

NGW

SSW

ESW/CCW

CGW

Colorado delta

Entrance to the

Gulf of California

PDW - Pacific Deep Water

PIW - Pacific Intermediate Water

NGW - Northen Gulf Water

SSW - Subequatorial Subsurface Water

ESW - Equatorial Surface Water

CGW - Central Gulf Water

CCW - California Current Water

3000 m

1000 m

500 m

250 m

50 m

(21)

10

Climate

The Gulf of California displays strong seasonality due to a monsoonal climate

driven by atmospheric changes in circulation, in particular the locations of the ITCZ, the

NPH and the SDL (Figure 1b). Ocean dynamics are also influenced by the seasonal

locations of the these weather systems [Douglas et al., 2007]. The wind associated with

the NPH and SDL act to control surface circulation and mixing in the Gulf, as well as

cause large evaporation rates [Pares-Sierra et al., 2003]. In the modern environment

sea-surface temperatures range from ~14 °C to over 30 °C [Roden, 1958; Bray, 1988a]. The

greatest seasonal fluctuation in sea-surface conditions occurs in the northern Gulf [Roden,

1958]. In part the large seasonality in climate is caused by an almost uninterrupted chain

of 1,000-3,000 m high mountain range on Baja California, resulting in a much reduced

oceanic influence and a more continental-type climate.

In winter (November to March) winds are northwesterly (Figure 1b) and come

from over the deserts of the American southwest. They are low in humidity acting to

increase evaporation rates, and are responsible for transporting large amounts of dust into

the northern and central Gulf [Douglas et al., 2007]. These winter winds also lower SST

and promote mixing in the upper ocean, causing the thermocline to disappear and

establishing upwelling. Upwelling begins along the eastern margin of the Gulf and may

expand across the basin [Thunell, 1998], acting to fuel high primary productivity. In

summer the northwesterly winds diminish and reverse, producing weak southerly winds

[Douglas et al., 2007]. The slackening of winds causes upwelling to relax and warm

Equatorial Pacific surface water is able to penetrate into the Gulf [Thunell, 1998]. In

summer and early fall (April to October) SST exceeds 29 °C due to increased solar

(22)

11

insolation, diminished upwelling, and the intrusion of warm Equatorial water. By the

middle of summer a thick layer (~150 m) of warm (>28 °C) water covers the central and

southern Gulf, creating a deep thermocline [Douglas et al., 2007]. The water column

becomes highly stratified and without the supply of nutrients from depth, the surface

water becomes depleted above the thermocline and primary productivity is reduced

[Thunell et al., 1996]. In addition, summer is the rainy season, with the northward

location of the ITCZ. Most of the precipitation occurs from July through September

[Calvert, 1966], and mainly falls on the south-eastern side of the Gulf on the Mexican

mainland [Douglas et al., 2007].

3. Material and Methods

The ~63.5 m long giant piston core MD02-2515 was collected from Guaymas

Basin (27° 29.01’ N, 112° 04.46’ W, 881 m water depth) in June 2002 during the second

leg of the MONA (Marges Ouest Nord Américaines) cruise of the RV Marion Dufresne

[Beaufort et al., 2002]. The age model used in this study was developed by Pichevin et al.

[2012] and is also described in McClymont et al. [2012]. Chronology of the section of

core analyzed in this study (top 48 m) is based on 17 accelerator mass spectrometry

(AMS)

14

C dates measured on bulk organic matter and two additional measurements on

planktonic foraminifera [Pichevin et al., 2012]. An age-depth diagram is shown in figure

3 for the 17 AMS

14

C dates used construct the age model.

The core is laminated for considerable portions of its length and lamination

thickness is variable [Cheshire et al., 2005]. Occasional intervals of non-laminated

sediment occur, some of which have been slightly to heavily bioturbated [Beaufort et al.,

(23)

12

2002; Figure 4]. The sediments contain varying amounts of biogenic and terrigenous

material, and range in composition from sand to silty clay to diatom ooze (Figure 4).

Sedimentation rates are very high, ranging from ~22 to 275 cm ka

-1

and averaging ~120

cm ka

-1

(Figure 4).

(24)

13

Figure 3. Age-depth diagram for the 17 AMS

14

C dates used to construct the age model

by Pichevin et al. [2012]. Interstadials are shown by shaded horizontal bars and

dinoflagellate cyst zones are separated by dashed lines. Heinrich events and the Younger

Dryas are highlighted in dark grey.

Age kyr BP 10 20 30 15 25 35

1a

1b

1c

1d

2a

2b

2c

3

!"#"$%&%

'()

*

+

,

-.

/

0

1&2%3425675#4

!8

!*

!+

!,

9:

0

20

40

60

Corrected depth

(m)

40

(25)

14

Figure 4. GISP II δ

18

O [Stuiver and Grootes, 2000], % total organic carbon (% TOC)

[Pichevin et al., 2012], % biogenic silica (% BioSi) [Pichevin et al., 2012], magnetic

susceptibility [Cheshire et al., 2005], sedimentation rate, lithology and structure

[Beaufort et al., 2002]. Bold curves are 10 point averages. Interstadials are shown by

shaded horizontal bars and dinoflagellate cyst zones are separated by dashed lines.

Heinrich events and the Younger Dryas are highlighted in dark grey.

Age kyr BP

% TOC

% BioSi

!

18

O (% )

GISP2

1a 1b 1c 1d 2a 2b 2c 3

2.0

2.5

3.0

15

25 35 45

-43 -41 -39 -37 -35 -33

!"#"$%&%

'()

*

+

,

-.

/

0

1&2%3425675#4

!8

!*

!+

!,

9:

Lithology Silty clay

Nannofossil silty clay Diatom-nannofossil silty clay Diatom silty clay

Diatom ooze Silty clayey diatom ooze Sand

Shell Shell fragment Fish debris Lamination Slump block or fold Slight bioturbation Heavy bioturbation Pyrite concretion Normal graded bedding Isolated layer

Sand lens or pocket Structure Legend

Lithology Struc

tur

e

10 20 30 15 25 35

Magnetic

susceptibility

Sedimentation

rate

0 100 200 300 -4 -2 0 2 4 S.I. units cm ka-1 10 20 30 15 25 35

(26)

15

3.1 Palynological Sample Preparation and Microscopy

One centimetre thick sediment slices were sampled every ~10 to 20 cm along the

core, corresponding to a sampling resolution of ~90 - 300 yrs. Additional intervals were

analyzed at a higher resolution to better constrain the duration of peaks in the

dinoflagellate cyst record. Each sample represents 4 to 45 years, averaging ~8 years of

accumulation. Recovery of dinoflagellate cysts and other palynomorphs from 286

samples was achieved using a standard palynological processing technique [Pospelova et

al., 2004]. Sediment samples of ~2.5 cm

3

were oven dried at 40°C and weighed

analytically. One tablet of exotic marker grain Lycopodium clavatum [Stockmarr, 1971;

Mertens et al., 2009] was added prior to sample treatment in order to determine total

palynomorph concentrations. The samples were treated with room temperature 10 % HCl

to remove carbonates, rinsed with distilled water, sieved through 120 µm mesh and

captured on a 15 µm mesh to remove both coarse and fine fractions. Siliceous material

was digested using room temperature 48-50 % HF for 2-3 days, followed by a second 10

% HCl treatment to remove precipitated fluorosilicates. Finally, the samples were rinsed

with distilled water, sieved, gently sonicated for up to 30 seconds and collected on a 15

µm mesh. One to two drops of the residue was mounted in glycerine jelly between a slide

and cover slip.

All palynomorphs were analyzed using a Nikon Eclipse 80i transmitting light

microscope at 600x and 1000x magnifications. Dinoflagellate cysts were identified

according to the paleontological taxonomy system described in Lentin and Williams

[1993]. Cysts of Polykrikos kofoidii and Polykrikos schwartzii follow the nomenclature of

(27)

16

possible, however due to morphological similarities and unfavourable orientations it was

not always possible, and some cyst taxa are grouped together. Brigantedinium spp.

consists of Brigantedinium cariacoense, Brigantedinium simplex, and other round brown

cysts, as archeopyles were not always observed due to unfavourable orientations or

folding. Selenopemphix quanta and cyst of Protoperidinium nudum are grouped together

as these two cyst species show a high degree of morphological similarity. Cyst type A has

been previously identified by studies in the Northeast Pacific [e.g. Radi et al., 2007;

Limoges et al., 2010]. Cyst type L is a spiny brown cyst and has been previously reported

by Price and Pospelova [2011] from Saanich Inlet, British Columbia (Canada). Other

spiny brown cysts of unknown affinity were placed in the spiny brown (SBC) category. A

list of all dinoflagellate cyst taxa recorded in this study and their corresponding biological

affinities is provided in Table 1. Plate pictures can be found on the following pages.

An average of 393 cysts were counted per sample (min. 298, max. 946).

Dinoflagellate cyst accumulation rates (cysts cm

-2

yr

-1

) were calculated by multiplying

the cyst concentration (cysts g

-1

), by the sedimentation rate (cm yr

-1

), and dry bulk

density (g cm

-3

). Dry bulk density measurements were determined by dividing the dry

sediment weight (g) by the wet volume (cm

3

).

(28)

17

Cyst species

(paleontological name)

Dinoflagellate theca

(biological name)

Autotrophic

Achomosphaera spp. ? Gonyaulax sp. Indet. Bitectodinium spongium unknown

Bitectodinium tepikiense Gonyaulax spinifera Impagidinium spp. Gonyaulax sp. Indet. Impagidinium aculeatum Gonyaulax sp. Indet. Impagidinium strialatum Gonyaulax sp. Indet. Impagidinium paradoxum Gonyaulax sp. Indet. Lingulodinium machaerophorum Lingulodinium polyedra Nematosphaeropsis labyrinthus Gonyaulax spinifera complex Operculodinium centrocarpum

sensu Wall & Dale 1966 Protoceratium reticulatum

Operculodinium aguinawense ?Protoceratium sp. Operculodinium israelianium ?Protoceratium sp. Pyxidinopsis reticulata Gonyaulacaceae undif. Polysphaeridium zoharyi Pyrodinium bahamense

Spiniferites belerius Gonyaulax scrippsae, G. spinifera complex Spiniferites ramosus Gonyaulax scrippsae, G. spinifera complex Spiniferites bentorii Gonyaulax digitalis, G. spinifera complex Spiniferites delicatus Gonyaulax spinifera complex

Spiniferites pachydermus Gonyaulax spinifera complex Spiniferites hyperacanthus Gonyaulax spinifera complex

Spiniferites bulloideus Gonyaulax scrippsae, G. spinifera complex Spiniferites mirabilis Gonyaulax spinifera complex

Spiniferites spp. Gonyaulax complex Spiniferites type 1. Gonyaulax complex Tectadodinium pellitum Gonyaulax spinifera complex Tuberculodinium vancampoae Pyrophacus steinii

--- Pentapharsodinium dalei

Heterotrophic

Dubridinium spp. Diplopsalid group

--- Gymnodinium spp.

--- Polykrikos kofoidii

--- Polykrikos schwartzii

--- Polykrikos cf. kofoidii

Brigantedinium spp. ?Protoperidinium spp. Brigantedinium cariacoense Protoperidinium avellana Brigantedinium simplex Protoperidinium conicoides Echinidinium aculeatum Protoperidinium sp. indet. Echinidinium delicatum Protoperidinium sp. indet. Echinidinium cf. zonneveldiae Protoperidinium sp. indet. Echinidinium granulatum Protoperidinium sp. indet. Echinidinium transparantum Protoperidinium sp. indet. Echinidinium spp. Protoperidinium sp. indet. Lejeunecysta oliva Protoperidinium sp. indet. Lejeunecysta sabrina Protoperidinium sp. indet. Lejeunecysta spp. Protoperidinium sp. indet.

--- Protoperidinium americanum

--- Protoperidinium nudum

--- sensu Wall and Dale 1968 Protoperidinium oblongum

Protoperidinium spp. Protoperidinium spp. indet. Quinquecuspis concreta Protoperidinium leonis Selenopemphix nephroides Protoperidinium sp. indet. Selenopemphix undulata Protoperidinium sp. indet. Selenopemphix quanta Protoperidinium conicum Stelladinium reidii Protoperidinium compressum Trinovantedinium applanatum Protoperidinium pentagonum Trinovantedinium variabile Protoperidinium sp. indet. Votadinium calvum Protoperidinium oblongum Votadinium spinosum Protoperidinium claudicans

Cyst type A unknown

Cyst type F unknown

Cyst type L unknown

Spiny brown cysts ?Protoperidinium sp. indet.

Table 1. Taxonomic citation of dinoflagellate cysts identified in this study. Thecate

equivalents are taken from Head [1996], Zonneveld [1997], Zonneveld and Jurkschat

[1999], Marret and Kim [2009], Matsuoka et al., [2009], and Verleye et al. [2011].

(29)

18

!"#$%&'

1

2

3

9

8

7

6

5

4

!"#$%&'(

)*+,-$./+%"0&1-2$23+4*2,*#1-5(&64#"%&7#*5&#*%&89&!3(

'(&!"#$%&'&(&)"*$+),-$.)":&;<+4&''.9=9:&5"+0%&'(

8(.>(&!"#$%&'&(&)"*#$/$'01)":&;<+4&9=.&?@A:&5"+0%&'(

B(.?(&!"#$%&'&(&)"*2#3$-/&+)":&;<+4&9=.?C@:&5"+0%&'(

=:&C(&!"#$%&'&(&)"*2./&$,$.)":&;<+4&9=.='8:&5"+0%&'(

@.A(&4-"$.02#3$-/0#2&2*,$56/&(.3)2:&;<+4&9=.='>:&5"+0%&'(

(30)

19

!"#$%&'

1

3

9

8

7

6

5

4

2

!"#$%&'

()*+,$-.*%"/&0,1$12*3)1+)#0,45&63#"%&7#)4&#)%&'8&!25

95&!"#$%&'("#"%)*)+,-+./'0-'/%):&#))1;&4,1;4&+)#<="%4&1<&$,%&$*0&1.&$,%&0)13%44:&>?*3&99-8@A:&4"*/%&95&

'5&BC4$&1.&1.#2+0-+/3'("#"%)*(+&.":&>?*3&8D-EFD:&4"*/%&'5&

@5&1'&430-+./"("%)*5'-+/4"&4,1;*<+&%0*3C4$#"&#)3,%10C"%:&>?*3&99-9G@:&4"*/%&95&

G-E5&60./,%&'("#"%)*,.#2/',+/0%)*4%<4=*H#""&#</&I#"%&9ADD&4,1;*<+&0)%3*<+="#)&#)3,%10C"%:&>?*3&

8D-D98&4"*/%&':&>?*3&8D-D'9:&4"*/%&95&

D5&60./,%&'("#"%)*,.#2/',+/0%)*J#)5*2/%#,+2%):&>?*3&8D-D98:&4"*/%&'5

F-K5&60./,%&'("#"%)*+$%"#+7.#3.:&#))1;&4,1;4&4$)*#$%/&7#4%&1.&0)13%44%4:&>?*3&8D-D@G:&4"*/%&95

A5&60./,%&'("#"%)&$C0%&95:&>?*3&8D-D@A:&4"*/%&95

(31)

20

!"#$%&'

1

2

3

9

8

7

6

5

4

10

11

12

(32)

21

!"#$%&'(&

)*+,-$./+%"0&1-2$23+4*2,*#1-5(&64#"%&7#*5&#*%&89&!3(

:(&!"#$#%&'#(&)*+&,&'#-);&<=+4&9>.>'?;&5"+0%&:(

8(&!"#$#%&'#(&)*+&$(.'##;&<=+4&9>.@>A;&5"+0%&:(

'(.B(&!"#$#%&'#(&)*+&$(.'##&C+$-&*%0D4%0&1*24%55%5;&<=+4&9>.>:9;&5"+0%&8(

@.>(&!"#$#%&'#(&)&$E1%&:;&<=+4&9>.@F>;&5"+0%&:(

F(&!"#$#%&'#(&)*+-,,.#/&-);&<=+4&9>.@>>;&5"+0%&8(

A(&!"#$#%&'#(&)*'01.)-);&<=+4&::.:'8;&5"+0%&:(

?(&!"#$#%&'#(&)*"0234/&'1-);&<=+4&::.9'?;&5"+0%&:(

:9(&!"#$#%&'#(&)*1#'0+#,#);&<=+4&::.9B9;&5"+0%&:(

::(.:8(&!"#$#%&'#(&)&4/(&34"&'020$(3-);&<=+4&9>.>:9.8;&5"+0%&8(

(33)

22

!"#$%&'

1

2

9

8

7

6

5

50 !m

3

4

!"#$%&'(

)*+,-$./+%"0&1-2$23+4*2,*#1-5(&64#"%&7#*5&#*%&89&!3&:;"%55&2$-%*<+5%0&;2$%0(

=(&!"#$%#&#'(")*+,-.')/"*+>&?@+4&9A.A=B>&5"+0%&=(

8(&0$%#&(")"*+,.$11"#*+>&?@+4&9A.CD=>&5"+0%&=(

E(&0*2$3%*1'(")"*+,4&)%&+.'&$>&?@+4&==.9CD>&5"+0%&=(

'(&!3"/&)#$(")"*+,-"+.1$5>&?@+4&==.9'C>&5"+0%&=(

C(&!3"/&)#$(")"*+,%&3"&%'$)-$>&?@+4&==.9CD>&5"+0%&=(

A(&!3"/&)#$(")"*+,511(>&?@+4&9A.AAC>&5"+0%&=(&

F(&6*23"(")"*+&511(&?@+4&9A.CA9>&5"+0%&=(

D(&GH5$&2/&73'#'.$3"(")"*+,&+$3"%&)*+>&?@+4&9A.A'B>&5"+0%&=(

B(&89+)'(")"*+&511(>&12"H,2;&+;&<-+$%&-+,-"+,-$5&$-%&*%$+4:"#$+2;>&?@+4&==.=E'>&5"+0%&=(

(34)

23

!"#$%&'

1

2

3

9

8

7

6

5

4

!"#$%&'(

)*+,-$./+%"0&1-2$23+4*2,*#1-5(&64#"%&7#*5&#*%&89&!3(

:(&!"#$%"&'"()#(*'+$',&-.;&<=+4&9>.>?:;&5"+0%&:(

8(&@A5$&2/&/,+-+)&,#0#$#"1*+23+$4"1*5%B5C&D#""&#B0&E#"%&:F>G;&<=+4&::.9H:;&5"+0%&:(

I(&5+-.0#$#"1*'.36"1;&<=+4&9>.>>';&5"+0%&:(

H(&5+-.0#$#"1*()#$+("1;&<=+4&9>.''G;&5"+0%&:(&

'(&@A5$&2/&/,+-+)&,#0#$#"1*$"0"1;&<=+4&::.9'F;&5"+0%&:(&

>(&7&3&$+)&1)8#9*%".$-.;&<=+4&9>.>'9;&5"+0%&:(

?(&7&3&$+)&1)8#9*$&)8,+#0&(;&<=+4&::.9H:;&5"+0%&:(&

G(&7&3&$+)&1)8#9*"$0"3.-.;&<=+4&9>.'>>;&5"+0%&:(

F(&7-&33.0#$#"1*,&#0##;&<=+4&::.9'?;&5"+0%&:(

(35)

24

!"#$%&'

1

2

3

9

8

7

6

5

4

!"#$%&'(

)*+,-$./+%"0&1-2$23+4*2,*#1-5(&64#"%&7#*5&#*%&89&!3(

:(&;<5$&2/&!"#$%&'%"()%"*"'+''&=>+4&9'.'?9@&5"+0%&:&

8(&;<5$&2/&!"#$%&'%"()(,-./&01''@&=>+4&9'.'::@&5"+0%&:(

A(&;<5$&2/&!"#$%&'%"()4/(&%"*"'+''@&=>+4&9'.BB8@&5"+0%&8(&

C(&2,-'3'+'3'45)/,4#6/045@&=>+4&9'.BCA@&5"+0%&:(&

B(&2,-'3'+'3'45)7&/34#/0458&=>+4&:CA.:@&5"+0%&:(

'()2,-'3'+'3'45)4/(&1"33696#+'/68&=>+4&9'.'CC@&5"+0%&:(

?(&;<5$&$<1%&D@&=>+4&::.9C8@&5"+0%&:(

E(&;<5$&$<1%&F@&=>+4&9'.'BA@&5"+0%&:(

G(&;<5$&$<1%&F@&=>+4&9'.'AE@&5"+0%&:(

(36)

25

3.2 Species Diversity and Multivariate Statistical Analyses

The Shannon-Wiener index, a diversity index, was calculated for each sample as

follows:

where H’ is the Shannon-Wiener index, S is the total number of species (species richness)

and p

i

is the relative abundance of species i. This diversity index is advantageous as it

takes into consideration the number of species present, as well as the “evenness” of the

species.

Statistical analyses were performed on dinoflagellate cyst relative abundances

using CANOCO 4.5 for Windows [ter Braak and Smilauer, 2002]. The data were

logarithmically [log(x+1)] transformed to increase the statistical weight of species that

occur in lower abundances, as these species are often characterized by a narrower

ecological affinity. Detrended Correspondence Analysis (DCA) was first used to test the

nature of variability within the dinoflagellate cyst assemblage. The length of the first

gradient in standard deviation units was 1.9, which indicates linear variation. Principal

components analysis (PCA) was performed, which reduces the dimensionality of the

dataset while maintaining the variation.

4. Results

4.1 Dinoflagellate Cyst Relative Abundances, Concentrations and Accumulation

Rates

Diverse, abundant and well preserved dinoflagellate cyst assemblages were

recorded from 286 samples in the upper 48 m of core MD02-2515. The most dominant

(37)

26

dinoflagellate cyst taxa found consistently throughout were Brigantedinium spp. (9-84

%). Other important contributors were Operculodinium centrocarpum sensu Wall and

Dale 1966 (0-80 %), Spiniferites pachydermus (0-81 %), S. bentorii (0-54 %),

Dubridinium spp. (0-43 %), Lingulodinium machaerophorum (0-33 %), Quinquecuspis

concreta (0-33 %), cyst of Polykrikos cf. kofoidii (0-30 %), and Echinidinium aculeatum

(0-10 %). A number of peaks are observed throughout the record where one taxon

increases in abundance by ~30-80 % from the preceding samples, and subsequently

declines by a similar magnitude in the following samples. These peaks are almost

exclusively produced by O. centrocarpum, S. pachydermus, and S. bentorii, all of which

are autotrophic taxa (Figure 5).

A total of 62 dinoflagellate taxa were identified (Table 1). The number of species

in each sample ranged from 13 to 33, with an average of 24. The Shannon-Wiener index

varied from 0.84 to 2.49, averaging 1.8 and was lowest from ~36-23 ka, and was highest

in the Holocene and at the bottom of the core ~38-39 ka (Figure 6). Dinoflagellate cyst

concentrations vary by two orders of magnitude, from ~3,500 cysts g

-1

to ~183,000 cysts

g

-1

, averaging ~18,500 cysts g

-1

. Large fluctuations are also apparent, particularly in cysts

of autotrophic taxa in sediments younger than ~22.8 ka (Figure 5). In general

concentrations of heterotrophic (H) taxa are much greater than concentrations of

autotrophic (A) taxa, and is reflected in the H/A ratio (Figure 5). Heinrich event 1 is one

notable exception to this trend.

Total dinoflagellate cyst accumulation rates vary from ~36 cysts cm

-2

yr

-1

to 8,150

cysts cm

-2

yr

-1

, averaging ~1,580 cysts cm

-2

yr

-1

. Cyst accumulation rates are highest from

22.3 to 16.9 ka, roughly corresponding to the LGM and HE1. During this interval cyst

(38)

27

accumulation rates show considerable fluctuations and a general increasing trend (Figure

6). The high cyst accumulation rate during this period is predominantly driven by an

increase in sedimentation rate from ~170 cm ka

-1

to ~280 cm ka

-1

as well as high

concentrations of cysts of heterotrophic taxa. In addition, this interval shows many abrupt

peaks in the record, which are almost exclusively the result of significant increases in the

production of autotrophic taxa. At ~16.9 ka when the sedimentation rate drops by ~50 %,

the total cyst accumulation rate declines significantly, followed by a slow decrease before

levelling out at ~15.1 ka (Figure 6). Although only relative abundances of individual

dinoflagellate cyst taxa are displayed in figure 5, cyst concentrations and accumulation

rates show similar trends.

(39)

28

Fig

ure

5.

Relative abundances (%) of selected

dinoflagellate cyst taxa. Dinoflagellate cysts species that characterize particular

zones are highlighted in color. Interstadials are shown by shaded horizontal bars and dinoflagellate cyst zones are separated by

dashed lines. Heinrich events and the Younger Dryas are highlighted in dark grey.

1a 1b 1c 1d 2a 2b 2c 3 10 20 30 Age kyr BP 0 10 20 30 Lingulodinium machaerophorum 0 10 Nematosphaeropis labyrinthus 0 10 20 30 40 50 Operculodinium centrocarpum 0 10 20 Spiniferites bentorii Spiniferites pachydermus Tuberculodinium vancampoae Brigantedinium spp. Protoperidinium spp. Dubrinidium spp. Echinidinium aculeatum Quinquecuspis concreta Cyst of Polykrikos schwartzii Cyst of Polykrikos cf. kofoidii Stelladinium reidii Cyst type A SBC Cyst type L 0 10 20 30 40 81 60 78 72 36 44 53 54 31 80 56 57 74 69 0 10 20 30 40 0 10 20 0 10 20 0 20 40 60 0 10 0 10 0 10 0 10 0 5 0 5 0 5 0 5 !"#"$%&% '() * + , - . / 0 1&2%3425675#4 !8 !* !+ !, 9: 15 25 35 % Zones

(40)

29

4.2 Statistical Analyses and Dinoflagellate Cyst Zones

Principal components analysis performed on logarithmically transformed relative

abundance data yields four axes explaining 21.4 %, 11.8 %, 9.6 % and 8.0 % of the

variance respectively, for a total of 50.7 % (Figure 6). It provides insight into similarities

and differences in the dinoflagellate cyst assemblages recorded throughout the core.

Figure 7, a PCA biplot, shows the ordination of samples and species along the two most

dominant ordination axes, PCA1 and PCA2. Most cysts of autotrophic species are

ordinated on the positive side of PCA1.

Based on sample scores from PCA axes 1, 2, 3, and 4, and visual inspection of the

dinoflagellate cyst abundances, three major dinoflagellate cyst zones were established

with seven subzones. Zone 1 (40 - 22.8 ka) is characterized by negative PCA1 and PCA2

values, whereas zone 2 (22.8 - 11.1 ka) is characterized by mostly positive PCA1 and/or

PCA 2 values (Figure 6). Zone 3 has the highest positive PCA 2 values in the record.

Division of the subzones also involved PCA3 and PCA4 (Figure 6). Although all zones

are dominated by Brigantedinium spp., zone 2 is characterized by the consistent presence

and often high and fluctuating values of Operculodinium centrocarpum (0.5-80 %,

average 10.4 %) (Figure 5). O. centrocarpum in zone 1 never constitutes more than 5 %

of the total assemblage. Samples in dinoflagellate cyst zones 1 and 3 predominately

cluster together compared to samples found in zone 2, which show greater variation

(Figure 7). Greater dissimilarities in zone 1 are however observed in PCA3 and PCA4

(Figure 6). Characteristic species of each dinoflagellate cyst subzone are highlighted in

color in Figure 5.

(41)

30

Figure

6

. Sedimentation rate; concentrations and fluxes for total, autotrophic and heterotrophic taxa; ratio of heterotrophic (H) to

autotrophic (A) taxa, and A/H; Shannon

-Wiener index with colored vertical lines highlighting distinct changes; and sample scores for

PCA axes 1

-4, where different colors highlight changes between zones. Interstadials are shown by shaded horizontal bars and

dinoflagellate cyst zones are separat

ed by dashed lines. Heinrich events and the Younger Dryas are highlighted in dark grey.

Age kyr BP 18.3 10.3 15.5 8.8

A/H

H/A

Shannon-W

iener

inde

x

-1

0

2

-2

-1

0

1

2

-1

0

1

2

-1

0

1

2

10 15 20 25 30 35

Total

dino

flagella

te

cy

sts

Aut

otr

ophic

taxa

Het

er

otr

ophic

taxa

1a 1b 1c 1d 2a 2b 2c 3 !"#"$%&% '() * + , - . / 0 1&2%3425675#4 !8 !* !+ !, 9: Conc en tr ations (shaded g rey) x 10 4 cy sts g -1 F lux es (black) x 10 3 cy sts cm -2 yr -1 0 2 4 6 0 2 4 6 0 2 4 6 0 2 4 0.6 1.1 1.6 2.1 2.6 0 2 4 6 8 0 2 4 6 0 2 4 0 20 40 PC A1 (21.4 %) PC A2 (11.8 %) PC A3 (9.6 %) PC A4 (8.0 %)

1

cm k yr -1 0 100 200

Sedimen

ta

tion

ra

te

(42)

31

Figure 7. Ordination diagram generated by principal components analysis (PCA)

showing dinoflagellate cyst taxa (black arrows) and ordination of samples (colored dots

and triangles). Abbreviations: Bspp, Brigantedinium spp; Proto, Protoperidinium spp;

Dubr, Dubrinidium spp.; CystA, Cyst type A; SBC, spiny brown cysts and Echinidinium

species; CystL, Cyst type L; Psch, cyst of Polykrikos schwartzii; Pkof, cyst of Polykrikos

kofoidii; Pcfk, cyst of Polykrikos cf. kofoidii; Ocent, Operculodinium centrocarpum

sensu Wall and Dale 1966; Sben, Spiniferites bentorii; Spac, Spiniferites pachydermus;

Nlab, Nematosphaeropsis labyrinthus; Lmac, Lingulodinium machaerophorum; Eacu,

Echinidinium aculeatum; Tvan, Tuberculodinium vancampoae; Stel, Stelladinium reidii;

Qcon, Quinquecuspis contreta; Operculodinium aguinawense; Pdal, cyst of

Pentapharsodinium dalei; Pzoh, Polysphaeridium zoharyi; Sram, includes Spiniferites

ramosus and Spiniferites bulloideus; Ssp1, Spiniferites type 1; Sspp, Spiniferites spp.;

Smir, Spiniferites mirablis; Squa, includes Selenopemphix quanta and cysts of

Protoperidinum nudum; Sneph, Selenopemphix nephroides; Sund, Selenopemphix

undulata; Imap; Impagidinium spp. Eigenvalues are shown at the bottom of the figure.

1.0

-1.0

1.0

DZ-1a DZ-1b DZ-1c DZ-1d DZ-2a DZ-2b DZ-2c DZ-3

Legend

Bspon

Imag

Lmac

Nlab

Ocent

Oagu

Pdal

Pzoh

Sram

Sben

Smir

Spach

Ssp1

Sspp

Tvan

Brig

Prot

Pam

Dubr

Eacu

Qcon

Psch

Pkoif

Pcfkoif

Sund

Sneph

Squan

Stell

Cysta

SBC

TypeL

!"#$%!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!! & ' ( ) *+,-.! /-01-23$ !415$2/-.6$%!!!!!!!!!!!!!!!!!!!!!!!78'&) 78&&9 787:; 7879 & !<6=6.-,1/$!>$03$2,-5$

!/-01-23$!+?!%>$31$%!@-,-!!!!!!'&8) ((8& )'8A B78A !C6=!+?!-..!$15$2/-.6$%!!!!!!

!!!!!!!!!!! !!!!!! !!!!!! &

PCA axis 1

PC

(43)

32

4.3 The Modern Analogue Technique

Reconstruction of sea-surface parameters using the modern analogue technique, a

transfer function method based on the similarity between samples, was attempted for this

core. Dinoflagellate cyst assemblages (relative abundances) from the core were compared

to a global reference dataset of 1429 assemblages from modern surface samples [dataset

from de Vernal, pers. com.] where sea-surface parameters are known, using the

procedures described in de Vernal et al. [2001]. Of the 286 core samples, 155 samples

had no analogues, thus quantitative reconstruction of sea-surface parameters was not

possible for this core.

5. Discussion

Guaymas Basin is very productive in the modern environment, as documented by

high diatom and biogenic silica fluxes in sediments traps [Thunell et al., 1994; Thunell,

1998], and by high dinoflagellate cyst concentrations in surface samples [Wall, 1986;

Martínez-Hernández and Hernández-Campos, 1991; this study]. Over the last 40 ka there

have been a number of important millennial scale climatic events, in particular D-O

interstadials and Heinrich events that have impacted both productivity and

ocean-atmosphere circulation in the Northeast Pacific. Here we demonstrate how dinoflagellate

cyst production has responded to changes in climate, oceanography, and productivity in

Guaymas Basin over the Late Quaternary and discuss the resulting implications for

interpretations of past climate change in the Northeast Pacific and the position of the

ITCZ.

(44)

33

5.1 The (Paleo-) productivity Signal

The Gulf of California is known for having high primary productivity in the

modern environment, with an average integrated primary productivity rate of

approximately 0.38 g C m

-2

day

-1

[Zeitzschel, 1969]. However rates of productivity can

be greater than 1 g C m

-2

day

-1

and may even exceed 4 g C m

-2

day

-1

[Douglas et al., 2007

and references within]. These rates are comparable to upwelling regions off the coast of

Baja California and North Africa, placing the Gulf of California among the world’s most

productive oceanic regions [Zeitzschel, 1969; Douglas et al., 2007]. High primary

productivity in the Gulf is recorded in surface sediment samples, which have high %

BioSi and high dinoflagellate cyst concentrations. In the northern Gulf dinoflagellate cyst

concentrations from surface sediments range from ~4,500 to ~17,000 cysts g

-1

[Pospelova

et al., 2008], while in La Paz (southwestern Gulf) concentrations range from ~200 to

27,000 cysts g

-1

[Kielt, 2006; Limoges et al., 2010]; and in Guaymas Basin one surface

sample had a concentration of 31,000 cysts g

-1

[this study]. Cyst concentrations from the

Gulf are in the same order of magnitude as other high productivity zones and upwelling

regions such as West Africa ∼130 - ∼65,600 cysts g

-1

[Bouimetarhan et al., 2009], the

Chilean continental margin ~525 - ~100,750 cysts g

-1

[Verleye and Louwye, 2010], and

the Northeast Pacific ~100 - 35,000 cysts g

-1

[Pospelova et al., 2008].

Dinoflagellate cyst concentrations and accumulation rates have fluctuated during

the Late Quaternary, but have remained relatively high throughout the core. Cyst

concentrations range from ~ 3,500 - 183,000 cysts g

-1

, averaging ~18,500 cysts g

-1

. Cyst

accumulation rates vary from ~36 - 8,200 cysts cm

-2

yr

-1

, averaging ~1,600 cysts cm

-2

yr

-1

. Although cyst concentrations and sedimentation rates (~120 cm ka

-l

) are relatively

Referenties

GERELATEERDE DOCUMENTEN

In the case of the camera exposed to isotropic microwave stray radiation in W7-X, power flux going through the aperture is first reduced to large extent.. However, multiple

In subsequent years, the theory was being developed starting with contractive Markov oper- ators in the works of Lasota, through non-expansive Markov operators in Szarek’s,, and

The references of Arab sources to moister than present-day climate in north Africa overlap with the 'early medieval warm era' of north-west Europe, and are seen as further support for

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of

De aangetroffen sporen leveren door hun recente datering of natuurlijke aard geen meerwaarde aan de archeologische kennis van de omgeving met uitzondering van de weg.

This researcher through reading news reports and editorial comments coupled with interviews with journalists who worked for the newspaper during the research period concluded that

In Chapter 2 of the dissertation, a total of 24 samples from sediment traps deployed in winter and summer monsoon seasons in the southwest Taiwan waters were analysed