• 沒有找到結果。

一萬六千五百年來東帝汶降雨變遷

N/A
N/A
Protected

Academic year: 2022

Share "一萬六千五百年來東帝汶降雨變遷"

Copied!
72
0
0

加載中.... (立即查看全文)

全文

(1)

୯ҥѠ᡼εᏢ౛ᏢଣӦ፦ࣽᏢ܌!

ᅺγፕЎ!

Department of Geosciences College of Science

National Taiwan University Master Thesis

!

΋࿤Ϥίϖԭԃٰܿࡆؙफ़ߘᡂᎂ!

Precipitation variability in East Timor during the Last 16,500 years

!

!

ഋਕѳ!

Jin-Ping Chen

!

ࡰᏤ௲௤;!؇οࢪ!റγ!

Advisor: Chuan-Chou Shen, Ph. D.

!

ύ๮҇୯ 102 ԃ 8 Д!

August, 2013

(2)
(3)

ᇞ ᇞᖴ

!

ૈ୼໩ճڗள೭ঁᏢՏǴाགᖴޑΓϼӭΑǶ२ӃࢂךޑࡰࡰᏤԴৣ؇οࢪ௲

௤ǴନΑाགᖴԴৣޑ஼ЈࡰᏤᆶЍ࡭ǴаϷкᅈऐЈޑፕЎǵੇൔঅׯǴ؇Դ

ৣჹܭϩ݋ϯᏢޑᆒዴᆶᏢೌ዗וΨ஥๏ך೚ӭ௴วǶ٠གᖴ᚟᚟୯ࡏ௲௤ǵԯԯݥ

ғ௲௤аϷ׵׵ਔߘԴৣ฻ӚՏα၂ہ঩Ǵ๏ϒޑࡌ᝼ᆶࡰ௲Ƕ᚟Դৣ܌ޑђғނ

ᏢǴаϷځдፐ୸ǵ཮᝼ύޑ૸ፕǴ፟ϒךჹܭђ਻ংǵђᕉნ׳ు΋ቫޑᇡ᛽ǹ

ᛥҥᏢߏаϷ݅݅ጬᏢߏ๏Αךॺ೭٤໿ೈ೚ӭԖҔޑගҢکࡰЇǴ੿ࢂኧΨኧό

మǶགᖴ݅݅ёᏢۊ஥ךΕߐႈ⇘ۓԃکჴᡍ࠻ᕴᕴಒ࿯ޑΕߐǴགᖴֆֆ۸ণᏢߏ

ᆶࠬࠬঅࢩԴৣᆶᔅךॺ௨ှΑ೚ӭሺᏔޑᅪᜤᚇੱǶགᖴԯԴৣکΐԀεᏢޑ࣍࣍

ഁᄆֻ௲௤ᄋඉගٮᅹ਼ӕՏનෳۓ೛ഢǴ٠གᖴླྀླྀ⃎ҏᏢۊǵቅቅЎฑᏢۊаϷ

මਥޕ㡷λۆᔅךෳۓΑεໆኬҁǶགᖴഋഋ܃ᆝᏢۊک໳໳ಹ৒ᏢۊޑፕЎǴаϷ

ڬྼ⣬Ꮲۊک஭஭ػᆢᏢߏޑᔅշǶΨᖴᖴךޑӕᏢษษ҅лکࡌࡌᏂлǴᗋԖךޑৎৎ

Γޑן࡭Ǵᡣךኖډ౥཰Ƕ!

অ᠐ᅺγ੤ޑ܌ᕇளޑᡏᡍǴځሽॶჹךٰᇥᇻຬၸЎᏧҁيǹךૈ୼ᇥǴ

ӧךڙֹӦ፦سᅺγ੤ޑ૽ግࡕǴό໻ჹܭӦ፦ᆶᏢೌԖΑ׳ు΋ቫᇡ᛽ǴΨԋ

ࣁΑ΋ঁ׳ӳޑΓǶ!

(4)

Abstract

Here we present replicated stalagmite analysis į18O records over the past 16.5

thousand years (ka, before 1950 AD) with a gap of 11.5-5.7 ka of three stalagmites from

Lekiraka cave in East Timor (8o47’10.8”S, 126o23’31.1”E; 626 m above sea level).

Lekiraka į18O record suggests a relatively dry condition during the glacial period,

comparing with the late Holocene. Depletion in18O during Younger Dryas (YD) stadial

suggests a wet condition with high precipitation, which is consistent with that of Liang

Luar cave record in Flores, but in contrary to a dry condition in the Asian monsoon (AM) territory. Insignificant į18O change during 16-12 ka indicates that a wet Bølling-Allerød

(BA) interstadial in AM territory is not clear at this low-latitude zone. We propose that

the intensified precipitation in YD and “muted” BA could be attributed to: (1) enhanced

Australian-Indonesian summer monsoon (AISM), (2) displacement of Intertropical

Convergence Zone (ITCZ), and/or (3) exposed Sunda shelf landmass lead to change of

the regional atmospheric circulation.

I

(5)

ᄔ ᄔा!

!

!!!!ҁࣴزճҔڗԾܿࡆؙญࢰ(ࠄጎΖࡋѤΜΎϩΜᗺΖࣾǴܿ࿶΋ԭΒΜϤࡋ

ΒΜΟϩΟΜ΋ᗺ΋ࣾǴੇܘϤԭΒΜϤϦЁ)ϐΟਲ਼ҡแǴख़ࡌ٠ϩ݋ 16.5 ίԃ

߻(ka, before 1950AD)аٰϐ୔ୱђНЎᡂᎂ(ܭ 11.5 Կ 5.7 ίԃڀԃж໔ᘐ)Ƕڀ

ख़౜܄ϐܿࡆؙҡแϐᛙۓ਼ӕՏન)į18O*ਔ໔ׇӈᡉҢǴ࣬ჹܭύఁӄཥШǴ၀

୔ୱϐ਻ংܭ҃ԛӇයೀܭ࣬ჹଳᔿϐރݩǶҁइᒵܭཥиζЕය(Younger Dryas,

YD)ਔ໔ࢤϐҡแ18O ࣬ჹ჆ЮǴᡉҢ၀ਔයࣁ࣬ჹዊᔸǵଯफ़ߘϐ਻ংރݩǴᆶ

ᎃ߈୔ୱ߻Γࣴز่݀΋ठ(Liang Luar cave, Flores)Ǵ٠ᆶڂࠠ٥ࢪۑ॥(Asian

Monsoon, AM)୔ୱђ਻ংइᒵܭ၀ਔය܌և౜ϐ࣬ჹଳᔿǵեफ़ߘރݩǴևϸӛ

ჹᔈǶҁइᒵύǴ27 ίԃԿ 23 ίԃ໔ϐίԃЁࡋ਻ং٣ҹݢ୏٠ό՟ڂࠠ٥ࢪۑ

॥୔܌ᡉҢϐܴᡉᆶமਗ਼ǴBølling-Allerød (B-A)ཪයҭ҂ԖܴᡉཪϯၞຝǴ౒ෳ

ёૈচӢࣁ;(΋)Ǵӑᐞহۑۑ॥(Australian-Indonesian summer monsoon, AISM)ϐ

ቚம<(Β)Ǵ໔዗஥ᒟӝ୔(Intertropical Convergence Zone, ITCZ)ϐՏ౽<аϷ0܈

(Ο)Ǵ൮дഌැ(Sunda Shelf)р៛܌Ꮴठϐ୔ୱН਻ൻᕉᡂᎂǶ!

II

(6)

Content

Abstract ...I

ᄔा ... II

Content ... III List of Figures ... V List of Tables ...VI

Chapter 1 Introduction ... 1

1.1 Inter-hemispheric anti-phasing behavior of millennium-scale climate events ... 1

1.2 Indo-Pacific warm pool ... 2

1.3 Australian-Indonesian monsoon ... 4

Chapter 2 Regional setting and methods ... 6

2.1 Study site and research material ... 6

2.1.1 Location of Lekiraka cave ... 6

2.1.2 Regional settings ... 8

2.2 Experiments ... 9

2.2.1 Subsampling ... 9

III

(7)

2.2.2 Labwares for U-Th dating chemical procedure ... 9

2.2.3 U-Th dating chemical procedure ... 9

2.2.4 U-Th dating Instrumentation ... 12

2.2.5 Stable oxygen isotope analysis Instrumentation ... 13

2.2.6 Hendy Test and Replication Test ... 14

Chapter 3 Results ... 15

3.1 U-Th dating results and age model ... 15

3.2 Oxygen and carbon stable isotope records ... 22

3.2.1 Hendy Test and Replication Test ... 22

3.2.2 Oxygen stable isotope time series ... 24

Chapter 4 Discussion ... 26

4.1 Fluctuation of hydroclimate in Australian-Indonesian monsoon territory since the last 16,500 years ... 26

4.2 Comparison with global and regional paleoclimate proxy records ... 33

4.2.1 Late Pleistocene --- From H1 to YD ... 33

4.2.2 Middle to Late Holocene ... 38

Chapter 5 Conclusions ... 40

IV

(8)

References ... 41

Appendix (I) U-Th concentration data and dating results ... 51

Appendix (II) į18O records ... 55

List of Figures

Figure 1-1. Reconstructed SST of equatorial Pacific sediment cores ... 3

Figure 1-2. Main atmospheric features in the modern western Pacific ... 5

Figure 2-1. Map with cave and marine sediment core locations ... 6

Figure 2-2. Stalagmite samples, from the cave to the laboratory ... 7

Figure 2-3. Average rainfall amount / temperature of Ossu ... 8

Figure 3-1. Plot of sample depth versus age for three selected stalagmites ... 16

Figure 3-2. Stalagmite MC110803-1 ... 18

Figure 3-3. Stalagmite MC110803-2 ... 20

Figure 3-4. Stalagmite 090721-2MC ... 21

Figure 3-5. Results of ‘Hendy Test’ of Lekiraka stalagmites ... 23

Figure 3-6į18O records of Lekiraka stalagmites ... 25

Figure 4-1. Map with site locations and moisture trajectories ... 27

Figure 4-2. Moisture trajectories for Laing Luar cave ... 28

Figure 4-3. Map with site locations and seasonal climate information ... 29

V

(9)

Figure 4-4. Present-day IPWP hydroclimate ... 30

Figure 4-5. Paleoclimate proxy records over the past 18,000 years ... 32

Figure 4-6. Simulation result of the global general circulation model ... 34

Figure 4-7. Reconstruction of Sunda shelf sea-level during at 12.31 ka BP ... 36

Figure 4-8. Simulated Indo-Pacific Walker circulation change in the LGM ... 37

Figure 4-9. Simulated LGM changes in the Indo-Pacific ... 37

Figure 4-10. Paleoclimate proxy records over the past 6,000 years ... 39

List of Tables

Table 3-1. 230Th age data of stalagmite MC110803-1 ... 17

Table 3-2. 230Th age data of stalagmite MC110803-2 ... 19

Table 3-3. 230Th age data of stalagmite 090721-2MC ... 21

VI

(10)

Chapter 1 Introduction

1.1 Inter-hemispheric anti-phasing behavior of millennium-scale climate events

High-resolution paleoclimate proxy records during the last glacial period have been

reported over past decades using different natural archives, such as ice cores (North

Greenland Ice Core Project members, 2004; Petit et al., 1999; Stenni et al., 2011),

stalagmites (Cruz et al., 2006; Dykoski et al., 2005; Fleitmann et al., 2004; Wang et al.,

2001; Wang et al., 2006), and marine foraminifera (Kiefer and Kienast, 2005; Mohtadi

et al., 2011; Xu et al., 2010). The anti-phasing variation between millennium-scale

climate changes at Heinrich event 1 (H1) stadial, Younger Dryas (YD) stadial, and

Bølling-Allerød (BA) interstadial during the last glacial-interglacial period over the

Northern Hemisphere (NH) and the Southern Hemisphere (SH) was revealed (Fig. 1-1)

(Barker et al., 2009). The phenomenon was suggested to be contributed to fluctuation of

high-latitude Atlantic meridional overturning circulation (AMOC), induced by rapid

fresh water discharge (Denton et al., 2010; McManus et al., 2004; Stenni et al., 2011).

However, AMOC reduction events and these millennium-scale climate events in

low-latitude proxy records are rather subtle since the last glacial maximum (LGM), and

only few terrestrial paleoclimate high-resolution proxy records were available at the

low-latitude zones (Cruz et al., 2009; Griffiths et al., 2009; Partin et al., 2007; Wang et

1

(11)

al., 2006). Therefore, in order to clarify the teleconnection between North Atlantic

dynamic and low-latitude atmospheric perturbation, further insight to past activity of

Pacific side low-latitude major climate system such as the Indo-Pacific warm pool

(IPWP) and the Australian-Indonesian monsoon (AIM), is crucial for deciphering global

climate change.

1.2 Indo-Pacific warm pool

Many marine sediment cores obtained from the IPWP had been studied (Kiefer and

Kienast, 2005; Visser, Thunell and Stott 2003; Xu et al., 2010). The paleoclimate proxy

records suggest a substantially different regional climate development, comparing to

those of the high-latitude NH. Instead of profound signature of millennial climate events,

most of the IPWP proxy records show a constant warming situation since the LGM

(Kiefer and Kienast, 2005; Rosenthal et al., 2003; Stott, 2002).

Previous studies indicate that the IPWP does not only act as a fundamental role in

the climate system of South-East Asia, but also be likely important for propagation and

amplification of millennial climate events (Abram et al., 2009; DiNezio et al., 2013;

Tokinaga et al., 2012).

2

(12)

Figure 1-1. Reconstructed SSTs of equatorial Pacific sediment cores compared with ice core į18O records. Comparison of equatorial Pacific Ocean sediment core records (modified from Rosenthal et al., 2003). (A) NGRIP ice core į18O record from Greenland (North Greenland Ice Core Project members, 2004). Planktonic foraminiferal derived (B) į18O and (C) SST record. (D) Surface sea water į18O record.

(E) Byrd ice core į18O record form Antarctica (Blunier and Brook, 2001).

MD97-2124 MD97-2124

SCS 18287-3 ODP 806B

TR163-19 VR21-30

3

(13)

1.3 Australian-Indonesian monsoon

Dynamics of AIM, baring strong connection with East Asian Monsoon (EAM)

system (Fig. 1-2), since the last glacial are rather less studied. Paleoclimate model

experiments (Wyrwoll et al., 2007), paleoproductivity data (Holbourn et al., 2005) and

pollen records (Kershaw, van der Kaars and Moss 2003) revealed that variation of SH

summer insolation influence Australian-Indonesian summer monsoon (AISM) rainfall

on orbital timescale. In contrast, Miller et al. (2005) suggested that the dominant AISM

forcing is the strength of the East Asian winter monsoon (Miller et al., 2005), which is

governed by the strength of air outflowing from the semi-permanent high-pressure

system over Siberia. During the Holocene epoch, the mechanisms that drive millennial-

and centennial- scale events is poorly known. The current available coral records

(Abram et al., 2007; Tudhope et al., 2001) and ocean sediment core records (Stott et al.,

2004; Visser, Thunell and Stott 2003) only provide limited information. Model

experiments simulated perturbation of regional atmospheric circulation and

millennial-scale events since the LGM (DiNezio et al., 2013; Tokinaga et al., 2012;

Zhang and Delworth, 2005), revealing significant influence of Sunda Shelf exposure on

AIM dynamics.

Previous studies have suggested that (1) the AIM system was dominated by

insolation intensity and/or sea-level on orbital time scale (Griffiths et al., 2009; Miller et

4

(14)

al., 2005), (2) exposure of the Sunda Shelf may profoundly affect regional climate, and

(DiNezio et al., 2013; Tokinaga et al., 2012) and (3) regional climate system was rather

heterogeneous than previous thought (DiNezio et al., 2013; Mohtadi et al., 2011).

Located at the modern southern border of Austral winter ITCZ, hydroclimate condition

of East Timor is dominantly affected by AM and AIM dynamics (An et al., 2000;

Griffiths et al., 2009). Therefore, by acquiring high-resolution proxy records from East

Timor stalagmites, our goal is to clarify how the AIM system responds to regional and

global changes, and have further insight on teleconnection between North Atlantic

perturbation and low-latitude climate system since the LGM.

Figure 1-2. Main atmospheric features in the modern western Pacific. Schematic view of major atmospheric system during austral winter (June to August) (a) and austral summer (December to Februarys) (b). The dominant flows in the middle and low troposphere are indicated by closed and open arrows, respectively (An et al., 2000).

a. b.

5

(15)

Chapter 2 Regional settings and methods

2.1 Study site and research material

2.1.1 Location of Lekiraka cave

Stalagmites were obtained from Lekiraka cave, Ossu, Viqueque, East Timor (Fig

2-2). This cave is located at 8o47’10.8”S, 126o23’31.1”E; 626 m above sea level, with

limestone cap layer about 20m. Lekiraka cave is about 650 km east to the Liang Luar

cave (Flores, Indonesia, 8o32’S, 120o26’E; 550 m above sea level) (Griffiths et al.,

2009).

Figure 2-1. Map showing cave and marine sediment core locations. Stars denote the locations of Lekiraka cave (yellow), Liang Luar cave (orange) (Griffiths et al., 2009), Gunung Buda (white) (Partin et al., 2007), and sediment core GeoB10053-7 (red) (Mohtadi et al., 2011).

6

(16)

Figure 2-2. Stalagmite samples, from the cave to the laboratory. (a) Stalagmites in the Lekiraka cave, highlighted by white circle. (b) Unprocessed stalagmite samples. (c) Stalagmite samples that were halved, polished, ready for subsampling.

a. b.

c.

7

(17)

2.1.2 Regional settings

The air temperature of Lekiraka cave is 25.3oC, with relative humidity of 97%

(August 3rd, 2011).The annual mean temperature of Ossu, the cave site is 26.1oC, with

mean annual rainfall of 1948 mm recorded by the Ossu weather station, for ~57%

falling during austral summer monsoon season (December–March) and ~17% falling

during austral winter monsoon season (June–September) (Fig. 2-3).

Figure 2-3. Average rainfall amount / temperature of Ossu. Long-term (1952-1974 AD) monthly average rainfall amount (grey bars) and temperature (black circles) of Ossu weather station. Data was obtained from Timor-Leste Meteorology Center.

8

(18)

2.2 Experiments

2.2.1 Subsampling

Stalagmite were halved and polished. For U-Th dating, powdered samples, 50-120

mg each, were milled along the growth layer. For stable oxygen and carbon isotope time series, powder samples, 50-100 Pg each, were milled along the growth axis. For ‘Hendy

Test’ (Hendy et al., 1971), powdered samples, 50-100 Pg each, were milled along the

coeval growth layers with interval of 1 mm. All sub-sampling procedure was performed

in class-10000 sub-sampling room and class-100 benches.

2.2.2 Labwares for U-Th dating chemical procedure

Labware used in this research included Teflon beakers, bottles, and columns;

PE/PP bottles, centrifuge tubes, vials, pipette tips, and porous PP column frits. In order

to reduce chemical blank during U-Th isotope measurement, all labwares were

acid-cleaned (Shen et al., 2003) prior to U-Th dating chemical procedure (Shen et al.,

2003; Shen et al., 2012).

2.2.3 U-Th dating chemical procedure

For reduction of potential containment, the chemical procedure for U-Th dating

was performed in a class-10,000 clean room with independent class-100 benches and

hoods. U-Th dating chemical procedure was developed in HISPEC, Department of

Geosciences, National Taiwan University (Shen et al., 2003).

9

(19)

Sample digestion

1. Stalagmite powdered sample was weighed and preserved in a 30-ml Teflon beaker.

2. Sample was covered with pure water, and dissolved gradually by adding appropriate

drops of 14N HNO3.

3. Appropriate amount of 229Th-233U-236U spike solution was added and weighed.

4. Ten drops of HClO4 were added for removal of organic matters.

5. One drop of FeCl2 was added for iron co-precipitation.

6. The sample solution was refluxed at 200oC on a hotplate over 12 hours to achieve

complete U-Th isotopic equilibrium and decomposition of organic matters.

7. The sample solution was dried at 250 oC.

Iron co-precipitation and centrifugation

1. Dehydrated sample was covered with 1-2 ml pure H2O, and dissolved with 1-2 ml 2

N HCl, then transferred from Teflon beaker into a PE centrifuge tube. The Teflon

beaker was cleaned with dilute aqua regia.

2. Appropriate amount of NH4OH was gradually added to the sample solution, until

Fe(OH)3 precipitate (for uranium and thorium co-precipitation).

3. Suspension was then separated by centrifugation and discarded, and residual was

rinsed by appropriate amount of pure H2O. This step was repeated three times.

4. Ten drops of 14 N HNO3 were added to dissolve the residual, and transferred back

10

(20)

into the original Teflon beaker.

5. Two drops of HClO4 were added to the sample solution and dried at 250 oC.

6. One drop of 14 N HNO3 was added to dissolve the dehydrated sample, and then dried

at 250 oC. This step was repeated for three times, to keep cations in nitric form.

7. Dehydrated sample was dissolved with 4-5 ml 7 N HNO3.

Uranium and thorium extraction by anion-exchange resin column

1. A 6-7 cm long column was packed with AG 1-X8 anion-exchange resin was prepared.

2. One column volume of pure H2O with one drop of 14 N HNO3 was added to remove

metal ion.

3. Repeatedly adding one drop of 7 N HNO3 was added three times. One column

volume of 7 N HNO3 was then added to condition the column.

4. Sample solution was loaded into the column.

5. Repeatedly adding one drop of 7 N HNO3 three times. One column volume of 7 N

HNO3 was then added to fully elute Fe3+.

6. The Teflon beaker was refluxed with 15ml 1% aqua regia and 0.005 N HF for 10

minutes and then rinsed with pure H2O.

7. Repeatedly adding one drop of 6 N HCl three times. One column volume of 6 N HCl

was then added. The thorium fraction was collected with the original cleaned 30-ml

Teflon beaker. Two drops of HClO4 were added to the beaker to decompose organic

11

(21)

matters.

8. Repeatedly adding one drop of pure H2O three times. One column volume of pure

H2O was then added. The uranium fraction was collected with a clean 10-ml Teflon

beaker. Two drops of HClO4 were added to the beaker to decompose organic

matters.

9. The two separated thorium and uranium fraction were dried at 250 oC for fully

dehydration.

10. One drop of HClO4 was added to dissolve dehydrated sample. The sample was then

dried at 250 oC to remove organic matter. This step was repeated for three times.

11. One drop of 14 N HNO3was added to dissolve dehydrated sample. The sample was

dried at 250 oC. This step was repeated for three times.

12. The dried thorium and uranium fractions were dissolved with 1% aqua regia and

0.005 N HF and then transferred into PE vial from Teflon beaker for instrumental

analysis.

2.2.4 U-Th dating Instrumentation

A Thermo-Fisher multiple-collector inductively coupled plasma mass spectrometer

(MC-ICP-MS), equipped with a secondary electron multiplier (SEM) and combined

with an Aridus sample introduction system was used for U-Th isotopic measurement.

U-Th dating techniques were developed in the High-Precision Mass Spectrometry and

12

(22)

Environment Change Laboratory (HISPEC), Department of Geosciences, National

Taiwan University (Shen et al., 2012). The instrument sensitivity is 2-4%, with

precision of ±1-2Å ı for abundance determinations of 50-200 fg 234U (1 – 4 ng 238U)

or 230Th. One international standard, New Brunswick Laboratories Certified Reference

Material 112A (NBL-112A) was used for calibrating mass fraction and intensity bias

(Shen et al., 2012).

2.2.5 Stable oxygen isotope analysis Instrumentation

Stable oxygen and carbon isotope analysis was performed in three laboratories.

The first is the Kano Laboratory of the Department of Environmental Changes, Kyushu

University. Isotopic measurement was conducted by Thermo-Finnigan Delta Plus mass

spectrometer with GASBENCH II gas separation system. The second is Stable Isotope

Laboratory of the Department of Earth Sciences, National Taiwan Normal University.

Isotopic measurement was conducted by Micromass IsoPrime isotope ratio mass

spectrometer (IRMS). The third is Stable Isotope Laboratory of the Department of

Geoscience, National Taiwan University. Isotopic measurement was conducted by

Thermo-Finnigan MAT 253 IRMS, with Kiel-II carbonate system. Oxygen and carbon

stable isotopic data were reported as į18O and į13C respectively, relative to the Vienna

Pee Dee Belemnite (VPDB) reference standard. The NBS-19 was used as calibration

standard (į18O = -2.20‰, į13C = +1.95‰). One-sigma external precision of the

13

(23)

measurements was better than 0.08‰ for į18O and 0.04‰ for į13C.

2.2.6 Hendy Test and Replication Test

‘Hendy Test’, a widely adapted testing method, was proposed by Hendy et. al.

(1971), In order to preclude kinetic effect dominant stalagmites from providing biased

paleoclimate proxy records. The Hendy test is based on two indicators that are

established by the į18O and į13C value of subsamples in coeval stalagmite growth layers.

First is the standard deviation of į18O value, and second is the covariation between į18O

and į13C values. If į18O within the same layers have small ı value (<0.5‰) and/or į18O

and į13C values are poorly correlated (R2<0.4), the stalagmite is considered to

precipitate without dominant kinetic fractionation, thus its į18O profile would be

considered a valid paleoclimate proxy. Another relatively new and widely accepted

testing method is ‘Replication Test’ (Dorale and Liu, 2004), by estimating the

consistency between contemporaneous į18O profiles of stalagmites in different position

of the cave. Dorale and Liu (2004) proposed that only when kinetic effect is absent or

affect the stalagmites in the exact same way, will the different stalagmite į18O profiles

be well duplicated, in which the latter case is fairly unlikely. This study adapted both

approaches, to provide a robust examination.

14

(24)

Chapter 3 Results

3.1 U-Th dating results and age model

After screening, three stalagmites, MC110803-1, MC110803-2, and 090721-2MC

were selected (Figs. 3-2 to 3-4). Age models of three selected stalagmite were

constructed by 28 dating points, Age (before 1950 AD) intervals are 16.58-0.39 ka for

MC110803-1, 14.62-0.87 ka for MC110803-2, and 13.16-11.62 ka for 090721-2MC

(Fig. 3-1, Table 3-1 to 3-3). However, there is still a time gap at 11.6-5.7 ka in the

spliced interval of 11.62 ka-5.64 ka. The average 238U concentration of these three

stalagmites is 285 ppb for MC110803-1, 266 ppb for MC110803-2, and 554 ppb for

090721-2MC, respectively. The detailed U-Th isotopic and concentration data and

dating results are available in Appendix (I). The average growth rates of three

stalagmites are 0.21, 0.092, and 0.056 mm/yr for MC110803-1, MC110803-2), and

090721-2MC, respectively.

15

(25)

Figure 3-1. Plot of sample depth versus age for three selected Lekiraka cave stalagmites. Green, blue, and orange lines denote stalagmite MC110803-1, MC110803-2, and 090721-2MC, respectively. Most of 2ı error bars are smaller than the symbols.

16

(26)

Table 3-1. 230Th age data of stalagmite MC110803-1.

Sample ID Depth(mm) Age (ka) (UURU ı

110803-1A-29 29 0.386 0.021

110803-1A-80 80 0.442 0.012

110803-1A-117 118 0.605 0.011

110803-1A-122 122 0.677 0.010

110803-1A-151 151 0.792 0.013

110803-1A-167 167 1.374 0.018

110803-1A-326.5 326.5 2.115 0.016 110803-1A-370.5 370.5 2.803 0.019 110803-1A-447.5 447.5 3.274 0.029 110803-1A-635.5 635.5 5.394 0.034 110803-1A-637.5 637.5 14.51 0.13

110803-1A-789 789 16.576 0.087

17

(27)

Figure 3-2. Stalagmite MC110803-1

10 cm

110803-1-29 386 Ʋ 21 yr

110803-1-4 1063 Ʋ 26 yr

110803-1-118 605 Ʋ 11 yr

110803-1-124 677 Ʋ 10 yr 110803-1-153

792 Ʋ 13 yr

110803-1-167 1374 Ʋ 18 yr 110803-1-316.5

2115 Ʋ 16 yr

110803-1-361 2803 Ʋ 19 yr 110803-1-366

3756 Ʋ 24 yr

110803-1-447.5 3274 Ʋ 29 yr 110803-1-635.5

5394 Ʋ 34 yr

110803-1-637.5 14509 Ʋ 134 yr 110803-1-660

17474 Ʋ 138 yr

110803-1-690 16652 Ʋ 91 yr 110803-1-740

13948 Ʋ 83 yr

110803-1-789 16576 Ʋ 87 yr 110803-1-800

15453 Ʋ 139 yr

Hiatus 110803-1-80 442 Ʋ 12 yr

18

(28)

Table 3-2. 230Th age data of stalagmite MC110803-2.

Sample ID Depth(mm) Age (ka) (UURU ı

110803-2A-10 10 0.874 0.014

110803-2A-142 142 1.769 0.033

110803-2A-180 180 2.062 0.025

110803-2A-240 240 2.687 0.018

110803-2A-365 365 3.425 0.023

110803-2A-415 415 3.918 0.034

110803-2A-441 441 4.383 0.022

110803-2A-524 524 5.181 0.032

110803-2A-544 544 5.64 0.032

110803-2A-550 550 11.887 0.054

110803-2A-583.5 583.5 13.004 0.061

110803-2A-650 650 14.621 0.074

19

(29)

110803-2-2.5 886 Ʋ 18 yr 110803-2-10

874 Ʋ 14 yr

110803-2-135 7786 Ʋ 83 yr 110803-2-142

1769 Ʋ 33 yr

110803-2-180 2062 Ʋ 25 yr 110803-2-184

3042 Ʋ 36 yr

110803-2-240 2687 Ʋ 18 yr 110803-2-329

3720 Ʋ 17 yr

110803-2-333 7244 Ʋ 42 yr 110803-2-365

3918 Ʋ 34 yr

110803-2-373.5 4138 Ʋ 20 yr 110803-2-415

3918 Ʋ 34 yr

110803-2-500 7367 Ʋ 67 yr

110803-2-524 5181 Ʋ 32 yr 110803-2-544

5640 Ʋ 32 yr

110803-2-550 11887 Ʋ 54 yr 110803-2-566.5

12985 Ʋ 67 yr

110803-2-583.5 13004 Ʋ 61 yr 110803-2-610

13235 Ʋ 74 yr

110803-2-630m 13328 Ʋ 64 yr 110803-2-640

14621 Ʋ 74 yr

110803-2-670 14295 Ʋ 80 yr 110803-2-673.5

13897 Ʋ 110 yr

10 cm

Hiatus 110803-2-441 4375 Ʋ 22 yr

Figure 3-3. Stalagmite MC110803-2

20

(30)

Table 3-3. 230Th age data of stalagmite 090721-2MC.

Figure 3-4. Stalagmite 090721-2MC

Sample ID Depth(mm) Age (ka) (UURU ı

110803-2A-3 3 11.618 0.064

110803-2A-40 40 12.496 0.061

110803-2A-60 60 12.774 0.062

110803-2A-71 71 13.16 0.20

090721-2MC-3 11618 Ʋ 64 yr

090721-2MC-14 15991 Ʋ 85 yr 090721-2MC-40

12496 Ʋ 61 yr

090721-2MC-60 12776 Ʋ 62 yr 090721-2MC-71

13159 Ʋ 200 yr

1 cm

21

(31)

3.2 Oxygen and carbon stable isotope records

3.2.1 Hendy Test and Replication Test

‘Hendy test’ (Hendy et al., 1971) was conducted along coeval growth layers, and all layers have less than 0.2‰ of į18O variation (1-sigma) (Fig. 3-5a), showing that

stalagmites deposited at an oxygen isotopic equilibrium condition. The R2for į18O and į13C value in coeval layers of three selected stalagmites MC110803-1, MC110803-2,

and 090721-2MC were depicted respectively by Figs. 3-5b to 3-5d. Several layersį18O

and į13C values show high covariance. However, various slopes of regression lines have little implications of dominant kinetic effect in the Lekiraka cave. Two

contemporaneous į18O records for MC110803-1 and MC110803-2 are well replicated

from 0.8 to 5.7ka (Fig. 3-6), combined with robust Hendy Test, this high replication

between records suggest that little kinetic fractionation and insignificant influence form

water-rock interaction.

22

(32)

Figure 3-5. Results of ‘Hendy Test’ of Lekiraka stalagmites. (a) Green, blue, and orange symbols denote the results of ‘Hendy Test’ of various stalagmite sections.

1-sigma value of all tested layers are less than 0.2‰, indicate a stable precipitation environment. (b)(c)(b) The linear regressions of subsamples from different stalagmites are subsequently color-coded. Various slopes of regression lines suggest remote possibility of dominant kinetic effect.

a.

b. c. d.

23

(33)

3.2.2 Oxygen stable isotope time series

Oxygen stable isotope time series of three selected stalagmites are given in Figure

3-7. į18O values of Lekiraka stalagmite records vary between -4.89‰ and -2.31‰

during 16.5-11.5 ka, -2.94‰ and -6.85‰ during 5.7-3.5 ka. A rapid drop of 1‰

occurred during 5.5 ka – 5.7 ka. This compiled record shows an 18O depletion tend from

late Pleistocene to Holocene (Fig 3-7).

Previous studies indicate that deglaciation process would essentially decrease the seawater į18O values for ~1‰ of decreasing (Dykoski et al., 2005; Schrag et al., 1996).

We assume that the shift of sea level is proportional to į18O change during deglaciation,

the reconstructed deglacial sea-level curve (Bard et al., 1996; Fairbanks et al., 1989;

Grant et al., 2012; Siddall et al., 2003) can be used to calculate decrease in į18O value

of ~0.5‰ during 16 to 11.5 ka, and additional ~0.5‰ decrease during the first half of Holocene (Griffiths et al., 2009). However, the fluctuation of moisture į18O values

during transformation of last deglaciation is difficult to be well estimated. The exposure

and flooding of Sunda Shelf could bring an unexpected effect on regional atmospheric

circulation and alternation of moisture amount and source to our study sites.

24

(34)

Figure 3-6. į18O records of Lekiraka stalagmites. Green, blue and orange line respectively denote į18O records of stalagmite MC110803-1, MC110803-2, and 090721-2MC. 230Th dates with 2ıġ error are color-coded with stalagmites. Vertical colorful bars indicate Heinrich event 1 (H1) stadial, Bølling-Allerød (B-A) interstadial, and Younger Dryas (YD) stadial defined in previous studies (Dykoski et al., 2005; Petit et al., 1999; Wang et al., 2001).

25

(35)

Chapter 4 Discussion

4.1 Fluctuation of hydroclimate in Australian-Indonesian monsoon territory since

the last 16,500 years

Modern (1952-1974 AD) climate data (Griffiths et al., 2009; National Directorate

of Meteorology and Geophysics, Timor-Leste), long-term (1981-2011) regional

climatology data (http://www.esrl.noaa.gov, Physical Science Division, Earth System

Research Laboratory, NOAA), and moisture-source trajectory modeling (Draxler and

Rolph, 2013) (Figs. 4-1 to 4-3) show the same hydrological condition in our Lekiraka

cave and published Liang Luar cave in Western Flores (Griffiths et al., 2009) (Figs. 4-1

to 4-2). The agreements indicate the rainfall change in Lekiraka and Liang Luar caves is

dominantly governed by same climatic condition under the Australian-Indonesian

summer monsoon (AISM). The minor offset between Lekiraka and Liang Luar records

could be attributed to different travel distance of moisture source and different

summer/winter moisture ratios (Figs. 4-3 and 4-4). Therefore, /HNLUDND į18O record is

highly consistent with that of Liang Luar (Griffiths et al., 2009), with similar peak at

Younger Dryas (YD) stadial, and relatively low į18O during the Holocene. However,

there is a prominent shift of 1‰ of į18O value during 5.5-5.7 ka, which could be

attributed to recrystallization process of the stalagmites.

26

(36)

Figure 4-1. Map showing site locations and moisture trajectories. Stars mark the cave locations of Lekiraka cave (yellow, this study), Liang Luar cave in Western Flores (orange) (Griffiths et al., 2009), Bukit Assam and Snail Shell caves in Gunung Buda National Park in Borneo (white) (Partin et al., 2007), and also a marine sediment core GeoB10053-7 in the eastern Indian Ocean (red) (Mohtadi et al., 2011). Solid lines represent five-year average (2007-2011 AD) moisture trajectories of Lekiraka cave that were constructed by the HYSLPIT moisture trajectory model (Draxler and Rolph, 2013).

Blue lines represent the moisture trajectories for rainy season (December – March), and red lines represent moisture trajectories for the other months from April to November.

27

(37)

Figure 4-2. Moisture trajectories for Laing Luar cave (orange star, Griffiths et al.,

2009). Stars mark the cave locations of Lekiraka cave (yellow, this study), Liang Luar

cave in Western Flores (orange) (Griffiths et al., 2009). Solid lines represent moisture

trajectories (September 2006 – April 2007) of Liang Luar cave that were constructed by

the HYSLPIT moisture trajectory model (Draxler and Rolph, 2013). Blue lines

represent the moisture trajectories for rainy season (December - March), and red lines

represent moisture trajectories for dry season (June - August) (Modified from Griffiths

et al., 2009).

28

(38)

Figure 4-3. Map showing site locations and seasonal climate information.

Remote-sensing SST data are averaged from January 2002 to February 2008 (Aqua MODIS, http://oceancolor.gsfc.nasa.gov). Stars mark the cave locations of Lekiraka cave (yellow, this study), Liang Luar cave in Western Flores (orange) (Griffiths et al., 2009), Bukit Assam and Snail Shell caves in Gunung Buda National Park in Borneo (white) (Partin et al., 2007), and also a marine sediment core GeoB10053-7 in the eastern Indian Ocean (red) (Mohtadi et al., 2011). Superimposed black arrows indicate the intensity and direction of scatterometer wind in (a) August and (b) February (Modified from Mohtadi et al., 2011).

a. August b. February

29

(39)

Figure 4-4. Present-day IPWP hydroclimate. Observed mean annual sea surface salinity (SSS) (Antonov et al., 2009). Solid black contour indicates the IPWP boundaries as defined by the annual mean sea surface temperature > 28ɗ (Reynolds, 2002). Dots denote the locations of marine proxy records used in DiNezio et al. (2013).

Bukit Assam & Snail Shell

GeoB 10053-7

Liang Laur Lekiraka

30

(40)

A prominent į18O shift of 1.8‰ between 16.5-11.5 ka and 5.7-0.4 ka in the

spliced Lekiraka record indicates a relatively dry glacial period and wet Holocene

in East Timor. This results generally correspond to the up-rising of eustatic sea

level since the last glacial maximum (Grant et al., 2012; Hanebuth, Stattegger and

Grootes 2000; Hanebuth et al., 2011; Sathiamurthy, 2006; Sidall et al., 2003).

However, in sediment records from core GeoB10053-7 (Mohtadi et al., 2011)

(Fig. 4-5), millennial-scale events left relatively clear signatures in its AISM

records, and the AISM strengthen rapidly in late Holocene. Lekiraka and Liang

Luar records have a constant of about 1‰ offset, which could be attributed to

different ratios of moisture (Fig. 4-1 to Fig. 4-4). The inconsistency between

ocean sediment records and stalagmite records could be attributed to complex

rainfall mechanism in this area (DiNezio et al., 2013; Griffiths et al., 2010;

Mohtadi et al., 2011). Compare Lekiraka and Liang Luar records to planktonic

foraminiferal Ʃ18O-inferred Australian-Indonesian winter monsoon (AIWM)

records of GeoB10053-7, highly consistent pattern could be observed (Fig. 4-5).

The anti-phasing relationship between GeoB10053-7 AIWM records and Liang Luar / Lekiraka į18O records could be attributed to the “pull-push” mechanism

between the AIM and the EAM (An et al., 2000; Rohling et al., 2009).

31

(41)

Figure 4-5. Paleoclimate proxy records over the past 18,000 years. D į18O record of NGRIP ice core, Greenland (North Greenland Ice Core Project members, 2004). (b) į18O record of Dongge cave, China (Dykoski et al., 2005). (c  į18O record of Liang Luar cave, Flores, Indonesia (Griffiths et al., 2009). (d į18O record of Lekiraka cave, East Timor (colored line). (e) (f) ǻ18O and Lithogenic/CaCO3 record of sediment core GeoB 10053-7, offshore Java (Mohtadi et al., 2011). (g) Sea-level reconstruction (Sidall et al., 2003). (h) Insolation intensity of 30oS (January).

32

(42)

4.2 Comparison with global and regional paleoclimate proxy records

4.2.1 Late Pleistocene --- From H1 to YD

A dry/cold YD condition shown by significant increase of į18O value in the Asian

monsoon (AM) territory stalagmite records (Dykoski et al., 2005; Hu et al., 2008; Jiang

et al., 2012; Wang et al., 2001) is absent in Lekiraka and Laing Luar records. Instead, a

slightly wet/warm YD condition could be observed in Lekiraka and Laing Luar records

(Griffiths et al., 2009), corresponded to intensification of the East Asian Winter

monsoon (EAWM) (Yancheva et al., 2007). Assuming monsoon moisture trajectories

were broadly similar to present time, a large proportion of summer monsoon trajectory

would have been occupied by land during late Pleistocene (Grant et al., 2012; Hanebuth

et al., 2011; Sathiamurthy, 2006; Sidall et al., 2003) in Maritime Continent region.

Limited moisture availability of source moisture greatly confined the development of

AISM, which suggests that southern displacement of the mean ITCZ position should be

responsible for increasing rainfall over research sites during YD, agreed with modeling

results (Dykoski et al., 2005; Griffiths et al., 2009; Yancheva et al., 2007; Zhang and

Delworth, 2005).

33

(43)

Figure 4-6. Simulation result of the fully coupled ocean-atmosphere global general circulation model (CM2.0) with the slab ocean model. Simulated annual mean precipitation anomaly (m/yr) base on thermohaline reduction scenario, a slightly positive precipitation anomaly can be observed over the Southern Indian Ocean and the Southern Pacific. Yellow star denote the location of Lekiraka cave (this study). The blue contour is the annual mean sea level pressure (SLP) anomaly with an interval of 0.4 hPa (Modified from Zhang and Delworth, 2005).

34

(44)

Millennium scale climate events such as YD, Heinrich stadial 1 (HS1) , and

Bølling-Allerød (B-A) interstadial, are ‘muted’ in Maritime Continent region records, in

contrast to NH low latitude paleoclimate records (Fig 4-5). Few evidences could be

found that the insolation intensity and sea-level change during H1 to B-A were

responsible for this phenomenon. Sea surface temperature (SST), a complementary

factor, should be involved. According to modeling results, reduced zonal tropical

Indo-Pacific Ocean SST gradient and/or Sunda Shelf exposure could induce slowdown

and/or displacement in the Walker circulation (Fig. 4-6 to Fig. 4-9) (DiNezio et al.,

2013; Sathiamurthy, 2006; Tokinaga et al., 2012; Zhang and Delworth, 2005). SST

records over the IPWP since the last glacial maximum (LGM) were reconstructed by

multiple ocean sediment cores, and constant warming since the LGM could be observed

throughout IPWP region (Kiefer and Kienast, 2005; Rosenthal et al., 2003; Stott et al.,

2004; Visser, Thunell and Stott 2003; Xu et al., 2010), suggesting that the SST

fluctuation, as well as eustatic sea-level change, could be critical to the past

precipitation history for Maritime Continent region.

35

(45)

Figure 4-7. Reconstruction of Sunda shelf eustatic sea-level during at 12.31 ka BP (Sathiamurthy, 2006). At 12.31 ka BP, Sunda Shelf was about 50m - 55m below present day sea-level (Modified from Sathiamurthy, 2006). Yellow, orange, white, and red stars respectively denote the locations of Lekiraka cave (this study), Laing Luar cave (Griffiths et al., 2009), Bukit Assam and Snail Shell cave (Partin et al., 2007), and sediment core GeoB 10053-7 (Mohtadi et al., 2011).

Bukit Assam &

Snail Shell

GeoB 10053-7

Liang Laur Lekiraka

36

(46)

Figure 4-8. Simulated Indo-Pacific Walker circulation change in the LGM. Changes in vertical velocity (Ȧ) over the equatorial Indo-Pacific simulated by HadCM3 in response to the LGM forcing (colors). Contours are annual-mean Ȧ simulated in the pre-industrial control experiment (DiNezio et al., 2013).

Figure 4-9. Simulated LGM changes in the Indo-Pacific. Change in precipitation during (a) austral summer (DJF) and (b) austral winter (JJA) simulated by HadCM3 in response to the LGM forcing. Yellow stars denote the location of Lekiraka cave (this study).The coastlines correspond to the 120 m isobaths of the present day ocean bathymetry (modified from DiNezio et al., 2013).

HadCM3 simulated rainfall and salinity change DJF and JJA

37

(47)

4.2.2 Middle to Late Holocene

The Sunda Shelf has been fully flooded since the early Holocene (Hanebuth,

Stattegger and Grootes 2000; Hanebuth et al., 2011; Sathiamurthy, 2006); thus, minor

eustatic sea-level fluctuation no longer dominated the regional hydroclimate. Little

evidence indicates that the ITCZ mean position or other cross equatorial climate forcing

was responsible for centennial to decadal perturbation in low-latitude paleoclimate

records (Fig 4-10). Stable regional IPWP SST during the Holocene (Visser, Thunell and

Stott 2003; Stott et al., 2004) implies that there could be an insignificant variation in the

adjacent terrestrial records. Two speculations attribute to the regional inconsistent

short-term perturbation: First, the AIM system has a complex and nonlinear respond to

regional climate forcing (Mohtadi et al., 2011; DiNezio et al., 2013); second, local

components such as environment or geographical characteristic of cave sites are rather

notable for paleoclimate proxy records in a relative short time perspective (Fairchild et

al., 2006; Lachiniet et al., 2009).

38

(48)

Figure 4-10. Paleoclimate proxy records over the past 6,000 years. D į18O record of Dongge cave, China (Dykoski et al., 2005). (b) Ti concentration record of Lake Huguang Maar, China (Yancheva et al., 2007). (c  į18O record of Lekiraka cave, East Timor (colored line). (d) į18O record of Liang Luar cave, Flores, Indonesia (Griffiths et al., 2009). (e) Ti concentration record of Cariaco Basin (Haug et al., 2001). (f) į18O record of Vostok ice core (Petit et al., 1999). Red lines denote the 5-point average of each proxy records respectively.

Weak ÅÅEASM ÆÆStrong

Weak ÅÅEAWMÆÆStrong South ÅÅITCZ ÆÆNorth

39

(49)

Chapter 5 Conclusions

1. Stalagmite į18O records and modern local hydroclimate condition of the Lekiraka

cave are highly consistent with that of the Liang Luar cave (Griffiths et al., 2009). It

indicates that East Timor and Flores have experienced the same climate variation in the

past. Agreement of stalagmite records between two sites also suggests that the Lekiraka

records can be interpreted as a proxy of the AISM over the last glacial-interglacial

period.

2. The evolution RI/HNLUDNDį18O records could be attributed to: (1) fluctuation of the

AISM activity, (2) migration of the ITCZ, and (3) eustatic sea-level change.

3. The dynamic and inter-connected behavior of the IPWP, the AM and AIM systems

since the LGM highlights the fundamental importance of the warm pool region for

understanding climate change throughout the tropics.

4. Inconsistency between marine and terrestrial paleoclimate records in AIM territory is

likely contributed to complex regional rainfall mechanism.

5. Detailed mechanism of how the Sunda Shelf exposure would affect the regional

atmospheric circulation is still unclear. More high-resolution terrestrial, marine proxy

records and model simulations are required.

40

(50)

Reference

Antonov, J. I. (2009) World Ocean Atlas, Salinity, 2, 184.

Abram, N. J., M. K. Gagan, Z. Liu, W. S. Hantoro, M. T. McCulloch & B. W.

Suwargadi (2007) Seasonal characteristics of the Indian Ocean Dipole during the

Holocene epoch. Nature, 445, 299-302.

Abram, N. J., H. V. McGregor, M. K. Gagan, W. S. Hantoro & B. W. Suwargadi (2009)

Oscillations in the southern extent of the Indo-Pacific warm pool during the

mid-Holocene. Quaternary Science Reviews, 28, 2794-2803.

An, Z. S. (2000) The history and variability of the East Asian paleomonsoon climate.

Quaternary Science Reviews, 19, 171-187.

Bard, E., B. Hamelin, M. Arnold, L. Montaggioni, G. Cabioch, G. Faure & F. Rougerie

(1996) Deglacial sea-level record from Tahiti corals and the timing of global

meltwater discharge. Nature, 382, 241-244.

Barker, S., P. Diz, M. J. Vautravers, J. Pike, G. Knorr, I. R. Hall & W. S. Broecker (2009)

Interhemispheric Atlantic seesaw response during the last deglaciation. Nature,

457, 1097-U50.

Blunier, T. & E. J. Brook (2001) Timing of millennial-Scale climate change in

Antarctica and Greenland during the Last Glacial Period. Science, 291, 109-112.

41

(51)

Cruz, F. W., S. J. Burns, I. Karmann, W. D. Sharp & M. Vuille (2006) Reconstruction of

regional atmospheric circulation features during the late Pleistocene in

subtropical Brazil from oxygen isotope composition of speleothems. Earth and

Planetary Science Letters, 248, 495-507.

Cruz, F. W., M. Vuille, S. J. Burns, X. F. Wang, H. Cheng, M. Werner, R. L. Edwards, I.

Karmann, A. S. Auler & H. Nguyen (2009) Orbitally driven east-west

antiphasing of South American precipitation. Nature Geoscience, 2, 210-214.

Denton, G. H., R. F. Anderson, J. R. Toggweiler, R. L. Edwards, J. M. Schaefer & A. E.

Putnam (2010) The Last Glacial Termination. Science, 328, 1652-1656.

DiNezio, P. N. & J. E. Tierney (2013) The effect of sea level on glacial Indo-Pacific

climate. Nature Geoscience, 6, 485-491.

Dorale, J. A. & Z. H. Liu (2009) Limitations of Hendy Test criteria in judging the

paleoclimatic suitability of speleothems and the need for replication. Journal of

Cave and Karst Studies, 71, 73-80.

Draxler, R. R. & Rolph (2013), G. D. HYSPLITHYbrid Single-Particle Lagrangian

Integrated Trajectory Model <http://www.arl.noaa.gov/ready/hysplit4.html>.

Dykoski, C. A., R. L. Edwards, H. Cheng, D. X. Yuan, Y. J. Cai, M. L. Zhang, Y. S. Lin,

J. M. Qing, Z. S. An & J. Revenaugh (2005) A high-resolution, absolute-dated

Holocene and deglacial Asian monsoon record from Dongge Cave, China. Earth

42

(52)

and Planetary Science Letters, 233, 71-86.

Fairbanks, R. G. (1989) A 17,000-year glacio-eustatic sea-level record - influence of

glacial melting rates on the Younger Dryas event and deep-ocean circulation.

Nature, 342, 637-642.

Fairchild, I. J., C. L. Smith, A. Baker, L. Fuller, C. Spotl, D. Mattey, F. McDermott &

Eimp (2006) Modification and preservation of environmental signals in

speleothems. Earth-Science Reviews, 75, 105-153.

Fleitmann, D., S. J. Burns, U. Neff, M. Mudelsee, A. Mangini & A. Matter (2004)

Palaeoclimatic interpretation of high-resolution oxygen isotope profiles derived

from annually laminated speleothems from Southern Oman. Quaternary

Science Reviews, 23, 935-945.

Grant, K. M., E. J. Rohling, M. Bar-Matthews, A. Ayalon, M. Medina-Elizalde, C. B.

Ramsey, C. Satow & A. P. Roberts (2012) Rapid coupling between ice volume

and polar temperature over the past 150,000 years. Nature, 491, 744-747.

Griffiths, M. L., R. N. Drysdale, M. K. Gagan, J. X. Zhao, L. K. Ayliffe, J. C. Hellstrom,

W. S. Hantoro, S. Frisia, Y. X. Feng, I. Cartwright, E. S. Pierre, M. J. Fischer &

B. W. Suwargadi (2009) Increasing Australian-Indonesian monsoon rainfall

linked to early Holocene sea-level rise. Nature Geoscience, 2, 636-639.

Hanebuth, T., K. Stattegger & P. M. Grootes (2000) Rapid flooding of the Sunda Shelf:

43

(53)

a late-glacial sea-level record. Science, 288, 1033-1035.

Hanebuth, T. J. J., H. K. Voris, Y. Yokoyama, Y. Saito & J. Okuno (2011) Formation and

fate of sedimentary depocentres on Southeast Asia's Sunda Shelf over the past

sea-level cycle and biogeographic implications. Earth-Science Reviews, 104,

92-110.

Haug, G. H., K. A. Hughen, D. M. Sigman, L. C. Peterson & U. Rohl (2001) Southward

migration of the intertropical convergence zone through the Holocene. Science,

293, 1304-1308.

Hendy, C. H. (1971) The isotopic geochemistry of speleothems—I. The calculation of

the effects of different modes of formation on the isotopic composition of

speleothems and their applicability as palaeoclimatic indicators. Geochimica et

Cosmochimica Acta, 35, 801-824.

Holbourn, A., W. Kuhnt, H. Kawamura, Z. M. Jian, P. Grootes, H. Erlenkeuser & J. Xu

(2005) Orbitally paced paleoproductivity variations in the Timor Sea and

Indonesian Throughflow variability during the last 460 kyr. Paleoceanography,

20.

Hu, C. Y., G. M. Henderson, J. H. Huang, S. Xie, Y. Sun & K. R. Johnson (2008)

Quantification of Holocene Asian monsoon rainfall from spatially separated

cave records. Earth and Planetary Science Letters, 266, 221-232.

44

(54)

Jaffey, A. H., K. F. Flynn, Glendeni.Le, W. C. Bentley & A. M. Essling (1971) Precision

measurement of half-Lives and specific activities of U-235 and U-238. Physical

Review C, 4, 1889-&.

Jiang, X. Y., Y. Q. He, C. C. Shen, X. G. Kong, Z. Z. Li & Y. W. Chang (2012)

Stalagmite-inferred Holocene precipitation in northern Guizhou Province, China,

and asynchronous termination of the Climatic Optimum in the Asian monsoon

territory. Chinese Science Bulletin, 57, 795-801.

Kershaw, A. P., S. van der Kaars & P. T. Moss (2003) Late Quaternary

Milankovitch-scale climatic change and variability and its impact on monsoonal

Australasia. Marine Geology, 201, 81-95.

Kiefer, T. & M. Kienast (2005) Patterns of deglacial warming in the Pacific Ocean: a

review with emphasis on the time interval of Heinrich event 1. Quaternary

Science Reviews, 24, 1063-1081.

Lachniet, M. S. (2009) Climatic and environmental controls on speleothem

oxygen-isotope values. Quaternary Science Reviews, 28, 412-432.

Linsley, B. K., Y. Rosenthal & D. W. Oppo (2010) Holocene evolution of the Indonesian

throughflow and the western Pacific warm pool. Nature Geoscience, 3,

578-583.

McManus, J. F., R. Francois, J. M. Gherardi, L. D. Keigwin & S. Brown-Leger (2004)

45

(55)

Collapse and rapid resumption of Atlantic meridional circulation linked to

deglacial climate changes. Nature, 428, 834-837.

Miller, G., J. Mangan, D. Pollard, S. Thompson, B. Felzer & J. Magee (2005)

Sensitivity of the Australian Monsoon to insolation and vegetation: Implications

for human impact on continental moisture balance. Geology, 33, 65-68.

Mohtadi, M., D. W. Oppo, S. Steinke, J. B. W. Stuut, R. De Pol-Holz, D. Hebbeln & A.

Luckge (2011) Glacial to Holocene swings of the Australian-Indonesian

monsoon. Nature Geoscience, 4, 540-544.

North Greenland Ice Core Project members. (2004) High-resolution record of Northern

Hemisphere climate extending into the last interglacial period. Nature, 431,

7005, 147-151.

Partin, J. W., K. M. Cobb, J. F. Adkins, B. Clark & D. P. Fernandez (2007) Stalagmite

records of tropical pacific climate since the last glacial maximum. Geochimica

Et Cosmochimica Acta, 71, A760-A760.

Petit, J. R., J. Jouzel, D. Raynaud, N. I. Barkov, J. M. Barnola, I. Basile, M. Bender, J.

Chappellaz, M. Davis, G. Delaygue, M. Delmotte, V. M. Kotlyakov, M. Legrand,

V. Y. Lipenkov, C. Lorius, L. Pepin, C. Ritz, E. Saltzman & M. Stievenard (1999)

Climate and atmospheric history of the past 420,000 years from the Vostok ice

core, Antarctica. Nature, 399, 429-436.

46

(56)

Reynolds, R. W., N. A. Rayner, T. M. Smith, D. C. Stokes & W. Q. Wang (2002) An

improved in situ and satellite SST analysis for climate. Journal of Climate, 15,

1609-1625.

Rohling, E. J., Q. S. Liu, A. P. Roberts, J. D. Stanford, S. O. Rasmussen, P. L. Langen &

M. Siddall (2009) Controls on the East Asian monsoon during the last glacial

cycle, based on comparison between Hulu Cave and polar ice-core records.

Quaternary Science Reviews, 28, 3291-3302.

Rosenthal, Y., D. W. Oppo & B. K. Linsley (2003) The amplitude and phasing of

climate change during the last deglaciation in the Sulu Sea, western equatorial

Pacific. Geophysical Research Letters, 30.

Sathiamurthy, E., H. K. Voris (2006) Maps of Holocene Sea Level Transgression and

Submerged Lakes on the Sunda Shelf. The Natural History Journal of

Chulalongkorn University, Supplement 2: 1-43

Schrag, D. P., G. Hampt & D. W. Murray (1996) Pore fluid constraints on the

temperature and oxygen isotopic composition of the glacial ocean. Science, 272,

1930-1932.

Shen, C.C., C.C. Wu, H. Cheng, R. Lawrence Edwards, Y.-T. Hsieh, S. Gallet, C.-C.

Chang, T.-Y. Li, D. D. Lam, A. Kano, M. Hori & C. Spötl (2012) High-precision

and high-resolution carbonate 230Th dating by MC-ICP-MS with SEM

47

(57)

protocols. Geochimica et Cosmochimica Acta, 99, 71-86.

Shen, C. C., H. Cheng, R. L. Edwards, S. B. Moran, H. N. Edmonds, J. A. Hoff & R. B.

Thomas (2003) Measurement of attogram quantities of Pa-231 in dissolved and

particulate fractions of seawater by isotope dilution thermal ionization mass

spectroscopy. Analytical Chemistry, 75, 1075-1079.

Siddall, M., E. J. Rohling, A. Almogi-Labin, C. Hemleben, D. Meischner, I. Schmelzer

& D. A. Smeed (2003) Sea-level fluctuations during the last glacial cycle.

Nature, 423, 853-858.

Stenni, B., D. Buiron, M. Frezzotti, S. Albani, C. Barbante, E. Bard, J. M. Barnola, M.

Baroni, M. Baumgartner, M. Bonazza, E. Capron, E. Castellano, J. Chappellaz,

B. Delmonte, S. Falourd, L. Genoni, P. Iacumin, J. Jouzel, S. Kipfstuhl, A.

Landais, B. Lemieux-Dudon, V. Maggi, V. Masson-Delmotte, C. Mazzola, B.

Minster, M. Montagnat, R. Mulvaney, B. Narcisi, H. Oerter, F. Parrenin, J. R.

Petit, C. Ritz, C. Scarchilli, A. Schilt, S. Schupbach, J. Schwander, E. Selmo, M.

Severi, T. F. Stocker & R. Udisti (2011) Expression of the bipolar see-saw in

Antarctic climate records during the last deglaciation. Nature Geoscience, 4,

46-49.

Stott, L. (2002) Super ENSO and global climate oscillations at millennial time scales

(vol 297, pg 222, 2002). Science, 298, 751-751.

48

(58)

Stott, L., K. Cannariato, R. Thunell, G. H. Haug, A. Koutavas & S. Lund (2004) Decline

of surface temperature and salinity in the western tropical Pacific Ocean in the

Holocene epoch. Nature, 431, 56-59.

Tokinaga, H., S. P. Xie, C. Deser, Y. Kosaka & Y. M. Okumura (2012) Slowdown of the

Walker circulation driven by tropical Indo-Pacific warming. Nature, 491, 439-+.

Tudhope, A. W., C. P. Chilcott, M. T. McCulloch, E. R. Cook, J. Chappell, R. M. Ellam,

D. W. Lea, J. M. Lough & G. B. Shimmield (2001) Variability in the El Nino -

southern oscillation through a glacial-interglacial cycle. Science, 291,

1511-1517.

Visser, K., R. Thunell & L. Stott (2003) Magnitude and timing of temperature change in

the Indo-Pacific warm pool during deglaciation. Nature, 421, 152-155.

Wang, S. Y. & Y. F. Qian (2001) Modeling of the 1998 East Asian summer monsoon by

a limited area model with P-sigma incorporated coordinate. Advances in

Atmospheric Sciences, 18, 209-224.

Wang, X. F., A. S. Auler, R. L. Edwards, H. Cheng, E. Ito & M. Solheid (2006)

Interhemispheric anti-phasing of rainfall during the last glacial period.

Quaternary Science Reviews, 25, 3391-3403.

Wang, Y. J., H. Cheng, R. L. Edwards, Z. S. An, J. Y. Wu, C. C. Shen & J. A. Dorale

(2001) A high-resolution absolute-dated late Pleistocene monsoon record from

49

(59)

Hulu Cave, China. Science, 294, 2345-2348.

Wyrwoll, K. H., Z. Y. S. Liu, G. Chen, J. E. Kutzbach & X. D. Liu (2007) Sensitivity of

the Australian summer monsoon to tilt and precession forcing. Quaternary

Science Reviews, 26, 3043-3057.

Xu, J. A., W. Kuhnt, A. Holbourn, M. Regenberg & N. Andersen (2010) Indo-Pacific

warm pool variability during the Holocene and Last Glacial Maximum.

Paleoceanography, 25.

Yancheva, G., N. R. Nowaczyk, J. Mingram, P. Dulski, G. Schettler, J. F. W. Negendank,

J. Q. Liu, D. M. Sigman, L. C. Peterson & G. H. Haug (2007) Influence of the

intertropical convergence zone on the East Asian monsoon. Nature, 445, 74-77.

Zhang, R. & T. L. Delworth (2005) Simulated tropical response to a substantial

weakening of the Atlantic thermohaline circulation. Journal of Climate, 18,

1853-1860.

50

數據

Figure 1-1. Reconstructed SSTs of equatorial Pacific sediment cores compared with  ice core į 18 O records
Figure 1-2. Main atmospheric features in the modern western Pacific. Schematic  view of major atmospheric system during austral winter (June to August) (a) and austral  summer (December to Februarys) (b)
Figure 2-1. Map showing cave and marine sediment core locations. Stars denote the  locations of Lekiraka cave (yellow), Liang Luar cave (orange) (Griffiths et al., 2009),  Gunung Buda (white) (Partin et al., 2007), and sediment core GeoB10053-7 (red)  (Moh
Figure 2-2. Stalagmite samples, from the cave to the laboratory. (a) Stalagmites in  the Lekiraka cave, highlighted by white circle
+7

參考文獻

相關文件

There’s Dave eating cake in a cave by the lake. There’s a cave by

一、生育功能的變遷 二、教育功能的變遷 三、經濟功能的變遷 四、保護功能的變遷 五、情感功能的變遷

In this class, we will learn Matlab and some algorithms which are the core of programming world. Zheng-Liang Lu 26

6 《中論·觀因緣品》,《佛藏要籍選刊》第 9 冊,上海古籍出版社 1994 年版,第 1

Teachers may consider the school’s aims and conditions or even the language environment to select the most appropriate approach according to students’ need and ability; or develop

(1) Western musical terms and names of composers commonly used in the teaching of Music are included in this glossary.. (2) The Western musical terms and names of composers

The accuracy of a linear relationship is also explored, and the results in this article examine the effect of test characteristics (e.g., item locations and discrimination) and

5.1.1 This chapter presents the views of businesses collected from the business survey, 12 including on the number of staff currently recruited or relocated or planned to recruit