<<

The University of Southern Mississippi The Aquila Digital Community

Dissertations

Summer 8-2007 A 3D REGIONAL MODEL OF THE INDONESIAN SEAS CIRCULATION Kieran Thomas Anthony O'Driscoll University of Southern Mississippi

Follow this and additional works at: https://aquila.usm.edu/dissertations Part of the Marine Biology Commons

Recommended Citation O'Driscoll, Kieran Thomas Anthony, "A 3D REGIONAL MODEL OF THE INDONESIAN SEAS CIRCULATION" (2007). Dissertations. 1284. https://aquila.usm.edu/dissertations/1284

This Dissertation is brought to you for free and open access by The Aquila Digital Community. It has been accepted for inclusion in Dissertations by an authorized administrator of The Aquila Digital Community. For more information, please contact [email protected]. The University of Southern Mississippi

A 3D REGIONAL MODEL OF THE INDONESIAN SEAS CIRCULATION

by

Kieran Thomas Anthony O’Driscoll

Abstract of a Dissertation Submitted to the Graduate Studies Office of The University of Southern Mississippi in Partial Fulfillment o f the Requirements for the Degree of Doctor of Philosophy

August 2007

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. COPYRIGHT BY

KIERAN THOMAS ANTHONY O’DRISCOLL

2007

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ABSTRACT

A 3D REGIONAL MODEL OF THE INDONESIAN SEAS CIRCULATION

by

Kieran Thomas Anthony O’Driscoll

August 2007

This study describes the ocean circulation of the Indonesian Seas based on results

using a 3D regional model. The study is divided into 3 parts. In the first part, Chapter

2, the basic properties o f a developed regional model o f the circulation o f the

Indonesian Seas are outlined. It is well known that the complex topography of the

region strongly influences temperature, salinity and current distributions there. One o f

the significant properties of this model is that all basic topographic features are

resolved. The model has four open ports to simulate inflow o f North Pacific Water

from the Mindanao Current, inflow o f South Pacific Water from the New Guinea

Coastal Current, outflow to the Pacific Ocean due to the North Equatorial Counter

Current, and outflow to the due to the . Total

transports through the open ports and port normal velocities are specified from

observations. Orlanski's conditions are employed at the open ports with port normal

velocity nudged to observed values and temperature and salinity to climatology. Port

channels are introduced so the effects of open boundary conditions do not impact the

dynamics of the main region. An additional friction was included in the vicinity of

some narrow passages and sills as a proxy for specific processes such as tides and

internal waves that occur within these topographic features. Four experiments are

discussed: seasonally varying and annual mean transports and port normal velocities

ii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. both with and without local winds. All experiments are totally spun up after 10 years.

This analysis uses data from the post spin up period only. The basic properties of

simulated total transports through the main passages in the region, surface circulation

and sea surface heights are discussed. The portion of North Pacific Water entering the

Indonesian Seas relative to that leaving through the North Equatorial Counter Current

port is fairly constant throughout the year. Most o f this water takes the western route

through the . The portion o f South Pacific Water entering the

Halmahera Sea compared to that exiting in the North Equatorial Counter Current

varies considerably with the seasons. Turning off the local winds does not

substantially influence the transport through main passages in the model domain.

Surface circulation patterns change substantially with the seasons. The role of

different terms in the heat and salt equations was investigated by dividing the region

into a number of boxes. For any given box, the sum of the horizontal advective fluxes

o f temperature (salinity) through all sides of the box is on the same order as the

vertical heat (salt) flux at the surface, interior nudging term, and the rate o f time

variation of the box integrated temperature (salinity). The comparison o f the basic

structure o f the model surface circulation, sea-surface heights and total transport

values through the main passages with observations appears satisfactory.

The main objective of the second part, Chapter 3, is to analyze the basic features of

potential temperature and salinity distributions in the Indonesian Seas simulated by

the model. The influence o f bottom topography on the formation o f temperature and

salinity distributions is considered by following the three major routes o f flow of

North Pacific and South Pacific Water through the Indonesian Seas. Major elevations

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. of bottom relief, such as the Sangihe Ridge; the topographic rise between the

Sulawesi Sea and Makassar Strait; the Dewakang Sill; the ridge between the Flores

and Banda Seas; the topographic rise between the Pacific Ocean and Morotai Basin;

the Lifamatola Sill; and the northern and southern Flalmahera Sills; break the region

down to separate basins having different temperature and salinity stratifications. The

differences in stratification are caused by these topographic features that act to

impede the advection of cold and salty water from a basin (located upstream) to the

neighboring basin (located downstream). Arguments are included to support this

conclusion. In the upper ocean (500m), the Indonesian Throughflow is primarily

shaped by North Pacific Water taking the route through the Makassar Strait. Deep

Banda Sea water is formed by the overflow of North Pacific Water across the

Lifamatola Passage into the . Below 500m South Pacific Water is blocked

by the Halmahera Sills and does not enter the Indonesian Seas. But in the upper ocean

(0-500m) SPW can probably penetrate into the Seram and Seas to mix with

NPW there.

There are no substantial structural changes o f potential temperature and salinity

distributions between seasons, though values o f some parameters o f temperature and

salinity distributions (e.g., magnitudes o f maxima and minima) can change. It is

shown that the main structure o f the observed distributions o f temperature and salinity

is satisfactorily displayed throughout the entire model domain. The calculated

transports of internal energy (heat) and salt mass through the Lombok and Ombai

Straits, and Passage in August and February are in reasonable agreement with

published observed and simulated data.

iv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. In the last part, Chapter 4, aspects o f turbulence and mixing in the Indonesian Seas

are presented and discussed. The results are based on the Mellor-Yamada 2.5

turbulence parameterization model. Though the importance of mixing in the

Indonesian Seas has been widely acknowledged, very few observations are available

and there have been no model studies of mixing or turbulent diffusion in the region.

The study is focused on turbulent diffusion and turbulent kinetic energy in the upper

mixed layer, the thermocline and in deep water near topographic features. Very large

turbulent kinetic energies and vertical turbulent diffusivities are seen around

topography and are important for the deep overflows found in the region. Large

turbulent energies and diffusivities found in the thermocline are important for the

diffusion o f temperature and salinity signatures found in the Indonesian Seas.

Monsoon winds and local currents lead to large diffusivities in the upper mixed layer.

v

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. to my mam and late dad

vi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ACKNOWLEDGMENTS

I would like to thank my PhD advisor Professor Vladimir Kamenkovich. Vladimir

has made a huge investment o f his time in me over the last 5 years for which I am

very grateful. I have certainly learned a lot about oceanography, and about life, from

Vladimir during this time and I hope our collaboration and friendship will continue

for many years to come. Thanks also to my committee members: to Professor Ralph

Goodman for his encouragement, enthusiasm for life and his friendship; to Professor

Dmitri Nechaev for his help with the model, the coding and many questions about

oceanography; to Drs. John Kindle and Pat Hogan from the Naval Research

Laboratory for their help and input into the process. I wish to thank the Naval

Oceanographic Office. NAVO have financed my PhD education while giving me

some time to do this work. Thanks also for making resources such as supercomputers

available to make the model runs. I also thank my colleagues at NAVO for their

continued encouragement and friendship. I thank everybody at USM/DMS for all

their help. Finally, thanks to my friends and family for all their love, friendship and

encouragement. I appreciate it.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. TABLE OF CONTENTS

ABSTRACT...... ii

DEDICATION...... vi

ACKNOWLEDGMENTS...... vii

LIST OF TABLES...... x

LIST OF ILLUSTRATIONS...... xi

LIST OF ABBREVIATIONS ...... xviii

CHAPTER

1. INTRODUCTION...... 1

1.1. Background ...... 2

1.2. O bjectives ...... 5 1.3. A Summary o f the Study ...... 5

2. MODEL DESCRIPTION AND PROPERTIES OF THE SIMULATED VELOCITY FIELDS...... 9

2.1. Introduction ...... 9 2.2. Model Description ...... 5 2.3. Results ...... 30 2.4. Estimate o f Terms in Heat and Salt Equations...... 48 2.5. Comparison with Observations ...... 56 2.6. Summary and Conclusions ...... 62

3. THE INFLUENCE OF BOTTOM TOPOGRAPHY ON TEMPERATURE AND SALINITY DISTRIBUTION ...... 69

viii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3.1. Introduction ...... 69 3.2. Transports of Mass, Internal Energy and S alt ...... 71 3.3. The Analysis of Temperature and Salinity Distributions ...... 12 3.4. Seasonal Variation ...... 96 3.5. Comparison with Observations ...... 107 3.6. Summary and Conclusions ...... 114

4. THE ANALYSIS OF DISTRIBUTIONS OF TURBULENCE CHARACTERISTICS BASED ON THE MODEL RESULTS...... 119

4.1. Introduction...... 119 4.2. The Turbulence Parameterization Scheme: The Mellor-Yamada (M-Y) 2.5 Model ...... 120 4.3. Results ...... 123 4.4. Discussion and Conclusions...... 143

5. SUMMARY...... 147

5.1. Future W ork ...... 149

APPENDICES...... 150

REFERENCES...... 158

ix

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF TABLES

Table

1. Subregions of the model domain where additional horizontal friction has been applied as follows: -r(i,j)u(i,j,k) and -r(i,j)v(i,j,k). The values of r are maximal at the centers o f the subregions and decay exponentially toward the boundaries ...20

2. List o f experiments. Transports out o f the model domain are positive; transports into the model are negative. In the first column 0 -, S-climatology, and wind stresses are characterized...... 29

3 . Estimates of different terms in the depth integrated heat, 0, and salt diffusion, S, equations for the box. August values are given for the seasonally varying experiment with local wind. Qs are the total transports through the sides o f the box. Negative values o f Q are into the box and positive values are out of the box. The last line gives the estimate of accuracy of calculations. Values for specific heat and reference density are, respectively, cp = 4 x 103 J °C'' kg'1, Po= 1025 kgnf3 ...... 54

4. The simulated transports o f mass, internal energy (heat), and salt through MC, NECC, and NGCC ports, along with Lombok and Ombai Straits, and Timor Passage for August and February. Additionally, the corresponding transports for the upper 500m are given. The specific heat capacity at constant pressure is cp=4x 103J°C'1kg'1. The mean density is given by p(rT025 kgm'3, 1 PW= 1015Js‘...... 72

C. The same as Table 3 but for different boxes in the region: C. l (above) the Banda Sea box, C.2 the Makassar Strait box and C.3 box to the north and east of ...... 155

x

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF ILLUSTRATIONS

Figure

2.1(a). Model bathymetry and domain (a) (above) and names o f important topographic features, open ports and islands (b). Note that in (b) the model domain is shown in x, y coordinates (i, j are given), and the depth in the grey region is less thanlOOm. Notice also that the island chain extending from Java in the west to Timor Island in the east is the Lesser Sunda Island Chain 15

2.1(b) ...... 16

2.2. Modeled mid-port normal velocity profiles at the 10 port (a), NGCC port (b), NECC port (c), and MC port (d). The profiles at (a), (c) and (d) are assumed time independent. Profiles for August and February are also given for (b). ...22

2.3. The evolution of Ev and Es with time. The characteristics Ev are shown in (a) for the seasonally varying experiment with local wind (exp 7; solid line) and the annual mean experiment with local wind (exp 1; broken line). The characteristics Es are shown in (b) for the same two experiments as (a). Variability of the integrated value of Ev in exp 1 is ~ 0. Ix l0 17kg m2/s2 and for Es is ~ 0 .5 x l0 13 kg m /s2 ...... 30

2.4(a). Transports across several sections of the model domain including the four open ports. Values for experiment 7 (seasonally varying with local wind) for August and February, and experiment 1 (annual mean with local wind) are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 (seasonally varying with no local wind) for August and February, and experiment 2 (annual mean with no local wind) are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration. Some arrows are slanted for graphic clarity...... 32

2.4(b) ...... 33

2.4(c) ...... 34

2.4(d) ...... 35

2.4(e) ...... 36

2.4(f) ...... 37

2.5. Transport time series through important passages and straits for four different model runs after 10 years o f model run. The seasonally varying experiments

xi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. with (exp 7) and without local winds (exp 8) are given by the solid and broken lines, respectively. The annual mean experiments with (exp 1) and without (exp 2) local wind are given by the . and x symbols, respectively. The sections shown are: (a) North Makassar Strait, (b) Lombok Strait, (c) Lifamatola Passage, (d) , (e) Obi-Seram passage, (f) Mangole-Buru passage, (g) Timor passage, (h) Ombai Strait. Note the different transport scales on the y-axes. For exps 7 and 8 (lines), transports are calculated every 10 days while for exps 1 and 2 transports are calculated only at times given by . and x symbols ...... 38

2.6(a). Surface velocities in m/s. Vectors for experiment 7 for August and February, and experiment 1 are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 for August and February, and experiment 2 are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration ...... 39

2.6(b) ...... 40

2.6(c) ...... 41

2.6(d )...... 43

2.6(e) ...... 44

2.6(f) ...... 45

2.7(a). Sea surface height. Values for experiment 7 for August and February, and experiment 1 are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 for August and February, and experiment 2 are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration ...... 47

2.7(b)...... 49

2.7(c)...... 50

2.7(d)...... 51

2.7(e)...... 52

2.7(f)...... 53

2.8. The eight boxes for which estimates o f terms in the heat and salt diffusion equations have been made. The boxes are: 1. Seram Sea, 2. Banda Sea, 3 Makassar Strait, 4. Lombok, 5. East Maluku Sea, 6. Karakelong-Sangihe- Morotai, 7. Halmahera Sea, and 8. West Sulawesi Sea ...... 55

xii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3.1. The locations of vertical sections AGHIJKF, ABCDEF, and LMND are shown by dots. Vertical profiles o f temperature and salinity are provided at points identified by letters A ,..., N. Sections, at which normal velocity and salinity are analyzed, are shown by solid lines. Numbers give August transports through these sections in Sverdrups (106 m3/s) in the direction of arrows. All arrows show transport directed normally to sections; slanting of some arrows is made for convenience o f plotting. Values o f indices i and j are given on the x- and y-axes. The Mindanao Current (MC), North Equatorial Counter Current (NECC), New Guinea Coastal Current (NGCC), and Indian Ocean (10) ports are seen where transports enter and leave the model domain (compare with Figure 2.1(b)) ...... 73

3.2(a). Vertical profiles o f potential temperature along the western route for August at points A, G, H, I, J, K, and F (a) (above) and the distribution o f potential temperature ort vertical section AGHIJKF (b) (shown in Figure 3.1), The bottom topography along this section is shown. Note the break of vertical scale at 300m. Locations o f points A, G, H, I, J, K, and F are indicated along the top o f the figure. Distance from point A is shown on the x-axis. Geographical names and locations are shown (see also Figure 2.1(b))...... 75

3.2(b) ...... 76

3.3(a). Vertical profiles of salinity along the western route for August at points A, G, H, I, J, K, and F (a) (above) and the distribution of salinity on vertical section AGHIJKF (b) shown in Figure3.1. Further explanations are similar to those in the caption to Figure 3.2...... 78

3.3(b) ...... 79

3.4. Isotachs of normal velocity (color) and isohalines (solid lines) at the Sangihe section (i=156; the values of j are indicated on the x-axis). Velocities are in m/s. For reference see Figure 3.1. The bottom topography is shown. The western route section (Figures 3.2(b), 3.3(b)) passes through this section at j=223 ...... 80

3.5. The same as Figure 3.4 but for the Dewakang section (j= l38; the values o f i are indicated on the x-axis). The vertical line shows the location of the section through the Dewakang sill (j=50) presented in Figure 8. The western route section (Figures 3.2(b), 3.3(b)) passes through this section at i=57 ...... 81

3.6. The same as Figure 3.4 but for the Flores-Banda section (i= 78; values of j are indicated on the x-axis). The western route section (Figures 3.2(b), 3.3(b)) passes through this section at j= l 10 ...... 8 2

xiii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3.7. v-velocities in m/s at the section through the Dewakang sill (i= 50; values of j are indicated on the x-axis). See Figure 3.5 for the location o f this section. The bottom topography is shown. The numbers on the x-axis (j) should be reversed with the largest on the left and the least on the right ...... 83

3.8. The distribution o f the POM 'vertical' velocity wpom on vertical section AGHIJKF (see Figure 3.1). Velocities are in m/s. The results are noisy but signs of wpom are provided at some points. Bottom topography along this section is shown. Distance from point A is shown on the x-axis. See Figure 3.3(b) for geographical names ...... 84

3.9(a). Vertical profiles o f potential temperature along the northern route for August at points A, B, C, D, E, and F (a) (above) and the distribution o f potential temperature on vertical section ABCDEF (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2...... 85

3.9(b)...... 86

3 .10. Vertical profiles o f salinity along the northern route for August at points A, B, C, D, E, and F (a) (above) and the distribution o f salinity on vertical section ABCDEF (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2 ...... 87

3.10(b)...... 88

3.11. The same as in Figure 3.8 but for the vertical section $ABCDEF$ (see Figure 3.1 ...... 89

3.12. The same as in Figure 3.4 but for the North Maluku section (j=187; values of i are indicated on the x-axis). The northern route section (Figures 3.9(b), 3.10(b)) passes through this section at i=151 ...... 90

3.13. The same as in Figure 3.4 but for the Lifamatola section ( p i 50; values o f i are indicated on the x-axis). The vertical line gives the location of the section through the Lifamatola passage (i=l 39) in Figure 3.14. The northern route section (Figures 3.9(b), 3.10(b)) passes through this section at i=137 ...... 91

3.14. v-velocities in m/s at the section through the Lifamatola passage (i= 139; values of j are indicated on the x-axis). See Figure 3.13 for section location. The bottom topography is shown. The numbers on the x-axis (j) should be reversed with the largest on the left and the least on the right...... 92

3.15(a). Vertical profiles of potential temperature along the eastern route for August at points L, M, N, and D (a) (above) and distribution o f potential temperature on vertical section LMND (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2 ...... 93

xiv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3.15(b) 94

3 .16(a). Vertical profiles o f potential temperature along the eastern route for August at points L, M, N, and D (a) (above) and distribution o f potential temperature on vertical section LMND (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2 ...... 95

3.16(b) ...... 96

3.17. The same as in Figure 3.4 but for the northern Halmahera sill ( p i 66; values of I are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=171 ...... 97

3.18. The same as in Figure 3.4 but for the Halmahera sea section ( p i 57; values of i are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=l 6 4 ...... 98

3.19. The same as in Figure 3.4 but for the southern Halmahera sill (p l4 9 ; values of i are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=156 ...... 99

3.20. The same as in Figure 3.8 but for the vertical section LMND (see Figure 3.1). Distances from point L are shown on the x-axis ...... 100

3.21. The same as in Figure 3.4 but for the section between Obi and Seram islands (i=148; values of j are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at j=144 ...... 101

3.22. The same as in Figure 3.4 but for the section between Obi and Seram islands (i=148; values of j are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at j= 144 ...... 102

3.23. The same as in Figure 3.4 but for the section across the Lombok Strait (j=l 10; values of i are indicated on the $x$-axis) ...... 103

3.24. The same as in Figure 3.4 but for the section across the Ombai Strait (i= 90; values of j are indicated on the $x$-axis) ...... 104

3.25. The same as in Figure 3.4 but for the section across the Timor Passage (i= 90; values of j are indicated on the $x$-axis)...... 105

4.1. Locations o f sections presented in the results. Section 1 - 2 is taken across the Northern Sangihe Ridge system from the Sulawesi Sea in the west to the Pacific Ocean in the east; section 3 - 4 is taken across the Dewakang Sill from the Southern Makassar Strait in the north to the in the south;

xv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. section 5 - 6 is taken across the Lifamatola Sill from the South Maluku Sea in the north to the North Banda Sea in the south; and section 7 - 8 is taken across the Halmahera Sea from the Obi-Seram passage in the south to the Pacific ocean in the north. The dots show the locations o f profiles o f 1, Sm and Sh. 124

4.2. Sections of (a) q2(m2/s2) (top) and (b) Kivi(m2/s) (bottom) across the North Sangihe Ridge. Locations of points 1 and 2 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 300 m ...... 128

4.2. (cont.). Section of (c) Kn(m2/s) (top) and profiles of (d) l(m) (bottom) across the North Sangihe Ridge, Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 300 m ...... 129

4.2. (cont.). Profiles of (e) Sm (top) and (f) Sh (bottom) across the North Sangihe Ridge. Locations of profiles are given by the dots in Figure 4.1 ...... 130

4.3. Sections of (a) q^(m2/s2) (top) and (b) K.M(m2/s) (bottom) across the Dewakang Sill. Locations of points 3 and 4 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100 m...... 131

4.3. (cont.). Section of (c) KH(m2/s) (top) and profiles of (d) l(m) (bottom) across the Dewakang Sill. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100m...... 132

4.3. (cont.). Profiles of (e) Sm (top) and (f) SH (bottom) across the Dewakang Sill. Locations of profiles are given by the dots in Figure 4.1 ...... 133

4.4. Sections of (a) q2(m2/s2) (top) and (b) Km(ih2/ s) (bottom) across the Lifamatola Sill. Locations o f points 5 and 6 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100 m...... 134

4.4. (cont.). Section of (c) KH(m2/s) (top) and profiles of (d) l(m) (bottom) across the Lifamatola Sill. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100 m...... 135

4.4. (cont.). Profiles of (e) SM (top) and (f) Sh (bottom) across the Lifamatola Sill. Locations of profiles are given by the dots in Figure 4.1 ...... 136

4.5. Sections of (a) q2(m2/s2) (top) and (b) KM(m2/s) (bottom) across the Halmahera Sea. Locations o f points 5 and 6 at the top o f the figures are given

xvi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100 m ...... 137

4.5. (cont.). Section of (c) Kn(m2/s) (top) and profiles of (d) l(m) (bottom) across the Halmahera Sea. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100 m...... 138

4.5. (cont.). Profiles of (e) SM (top) and (f) SH (bottom) across the Halmahera Sea. Locations of profiles are given by the dots in Figure 4.1 ...... 139

4.6. Maps of (a) q2(m2/s2) at the surface with climatological wind (top) and (b) no wind...... 140

4.6. (cont.) Maps of (c) q (m./s ) at 20m with climatological wind (top) and (d) no wind...... 141

4.6. (cont.) Maps of (e) q2(m2/s2) at 50m with climatological wind (top) and (f) no wind...... 142

A.l Some important features o f the smoothed model bathymetry: (a) Sangihe Ridge with north and south sills; (b) South Makassar Strait with the Dewakang Sill in the east and western passage near Kalimantan; (c) Lombok Strait;(d) Ombai Strait; (e) Timor Passage; (f) Halmahera Sea with north and south sills; (g) Seram Sea area with Lifamatola Passage to the northwest; Buru-Seram passage to the south; Obi-Halmahera and Obi-Seram passages to the east; and Mangole-Buru passage to the west; (h) Irian Jaya - Seram Passage; and (i) Labani Channel. Isobaths are in 1000's o f meters...... 152

xvii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF ABBREVIATIONS

ACCS Antarctic Circumpolar Current System

ADCP Acoustic Doppler Current Profiler

ARLINDO Arus Lintas Indonen

BBL Bottom Boundary Layer

BL Boundary Layer

COADS Comprehensive Ocean-Atmosphere Data Set

CTD Conductivity Temperature Depth

CTDO2 Conductivity Temperature Depth Oxygen

ECHO European Centre Hamburg Model

ECMWF European Center for Medium-Range Weather Forecasts

ET0P05 Earth Topography 5km

EUC Equatorial Undercurrent

GFDLMOM2 Geophysical Fluid Dynamics Laboratory Modular Ocean Model2

IES Inverted Echo Sounders

10 Indian Ocean

ITF Indonesian Throughflow

JADE Java Australia Dynamics Experiment

LANL Los Alamos National Laboratory

LLWBC Low Latitude Western Boundary Current

MC Mindanao Current

MIT GCM Massachusetts Institute of Technology General Circulation Model

M-Y Mellor-Yamada

xviii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NADW North Atlantic Deep Water

NEC North Equatorial Current

NECC North Equatorial Counter Current

NGCC New Guinea Coastal Current

NGCUC New Guinea Coastal Undercurrent

NICU New Ireland Coastal Undercurrent

NORDA Naval Ocean Research and Development Activity

NPW North Pacific Water

NWM North West Monsoon

OGCM Ocean Global Circulation Model

OZPOM Australian Princeton Ocean Model

POM Princeton Ocean Model

POM2K Princeton Ocean Model 2000 code

POP Parallel Ocean Program

PIES Pressure Inverted Echo Sounders

SEC South Equatorial Current

SEM South East Monsoon

SLA Sea Level Anomaly

SPGA Shallow Pressure Gauge Array

SPW South Pacific Water

SSH Sea Surface Height

STD Salinity Temperature Depth

TKE Turbulent Kinetic Energy

xix

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. TOPEX POSEIDON/ERS Typhoon Operational Experiment

UML Upper Mixed Layer

WEPOCS Western Equatorial Pacific Ocean Circulation Study

WOCE World Ocean Circulation Experiment

xx

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 1

CHAPTER 1

INTRODUCTION

The circulation of the Indonesian Seas is studied with a high resolution regional

model based on the Princeton Ocean Model (POM) (see Mellor 2004). The model

has been developed as a research tool to investigate important physical processes

occurring in the Indonesian Throughflow (ITF) and the different flow paths taken

by the ITF. Though previous modeling studies have made important contributions

to our understanding of the ITF, most have been concerned with transports of mass,

heat and salt through the Indonesian Seas and have not focused in detail on the

processes causing the transformation of water masses or paths taken by the water

masses within the Indonesian Seas region. In addition, the small number of observa­

tions in the region also limits our understanding of the ITF.

The Indonesian Seas are situated at low latitudes between the Indian and Pacific

Oceans. They are unique in that they form the only connection between ocean

basins at low latitudes at which interocean exchange can occur. As a result, the

ITF plays a vital role in earth‘s equatorial climate system. For example, the models

of Hirst & Godfrey (1.993), Verschell et al. (1995) and Murtugudde et al. (1998)

show that Sea Surface Temperature (SST) in the Pacific and Indian Oceans and

upper-layer heat storage are dependent on the ITF. The ITF is also thought to

play an important role in the global climate system, see, e.g., Gordon (2001) and

references therein. It is widely accepted that the ITF is strongly connected with

the global ocean thermohaline circulation cell (’’conveyor belt circulation system”)

whereby North Atlantic surface water sinks at high latitudes to form North Atlantic

Deep Water (NADW), spreads to the Indian and Pacific Oceans via the Antarctic

Circumpolar Current System (ACCS) and ultimately returns as surface water with

a major portion passing through the Indonesian Seas (see Gordon (1986), Broecker

(1991), Schmitz (1996 a, b)). Indeed, Shriver & Hurlburt (1997) and Goodman

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2

(1998) find there are profound effects on the ITF if the generation of NADW is

shutdown.

ITF water, that is water entering the Indian Ocean from the Indonesian Seas,

bears little resemblance to the water masses entering the Indonesian Seas at the Pa­

cific. These Pacific Ocean source waters are transformed while crossing the Indone­

sian Seas. The processes causing this transformation are not precisely known. The

topography of the Indonesian Seas region is highly irregular with the major basins

separated from each other by topographic features such as sills and narrow pas­

sages. A monsoon climate system is active in the region with a southeast monsoon

coming from Australia in the northern summer and a northwest monsoon coming

from Southeast Asia in the southern summer. The monsoons help dictate transport

through the ITF and flow paths taken through the region. Vertical diffusion of tem­

perature and salinity signatures is thought to be important in the thermocline and

near topographic features. Seasonal rain and river runoff from the islands add to the

complexity of the region.

1.1. Background

Observational and numerical studies show there is a transport of mass, heat and

salt through the Indonesian Seas in the direction of the Indian Ocean. Stammer et

al.(2Q03), using the MIT GCM that was fully constrained by WOCE data, found a

mean heat transport of about 1.1 PIT from the Pacific into the Indian Ocean. Piola

& Gordon (1984), using simple box models, found that a transport of 13.7 Sv from

the Indonesian Throughflow into the Indian Ocean with an average salinity of 33.55

was required to balance the heat content of the Indian Ocean. Schneider & Bar­

nett (1997), using a coupled ocean-atmosphere model (ECHO), calculated a mean

baroclinic component of heat transport from the western Pacific into the Indonesian

seas of 0.3PIT and a barotropic component of 0.4PIT. The seasonal heat transport

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3

toward the Indian Ocean is maximal at 1.4PIT in July and minimal at 0.1 PW in

February with fluxes referenced to 3A°C. Ganachaud & Wunsch (2000) used hy­

drographic sections, current meters with climatological wind fields, biogeochemical

balances and improved a priori error estimates in an inverse model, to improve es­

timates of the global circulation and heat fluxes. They found that a heat flux of

1.4PJT enters the Indian Ocean from the ITF. They also note that the hydrographic

data used in the model at this location came from the Java Australia Dynamics Ex­

periment (JADE) program in August 1989 only. Schiller et al. (1998), using a global

ocean circulation model (OGCM) with enhanced tropical resolution, found a mean

heat transport across the 1X1 transect (between Java and NW Australia) towards

the Indian ocean of 1.15PIT with a maximum of 1.86 PW in August and a minimum

of 0.58PIT in March. Gordon & McClean (1999) modeled the regional circulation

of the Indonesian Seas using the Los Alamos National Laboratory (LANL) Parallel

Ocean Program (POP) forced by ECMWF wind stresses for the period 1985 through

1995. They found a heat transport toward the Indian Ocean of 0.79PJT in August

1993 and 0.06PJT in February 1994. They found salt flux of 2.51 x 10 skg/s for

August 1993 and 3.0 x 10 7kg/s for February 1994. Hirst & Godfrey (1993), using

the Cox version of the Bryan-Cox OGCM, found a heat transport of 0.62PIT from

the Pacific Ocean toward the Indian Ocean. Lee et al. (2002) compared solutions

of a near-global circulation model with open and closed Indonesian passages from

1981 to 1997. The open Indonesian passage provided a heat flux of approximately

0.6P1T to the Indian Ocean. Godfrey (1996) in his review paper suggested that

a mean throughflow of lOS'u would transport about 0.5PJF of heat into the In­

dian Ocean. Ffield et al. (2000), using measurements of temperature and transport

from instruments deployed in the Labani Channel, calculated an observed heat flux

of 0.50P1T through the Makassar Strait for 1997. This included measurements of

0.63P1T during the La Nina months of December 1996 through February 1997 and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 4

0.39PW during the El Nino months of December 1997 to February 1998. Vranes et

al. (2002) calculated the heat transport through the Makassar Strait for the period

December 1996 - July 1998 using four different observed temperature profiles. Their

mean calculations for the period were 0.72 PW, 0.58PIT, 0.34 PW, and 0.53PIT.

Estimates of ITF transport from observational, numerical and other studies vary

from about 0 — 30ST toward the Indian Ocean, see, e.g., Wyrtki (1961), Godfrey &

Golding (1981), Murray & Arief (1988), Godfrey (1989, 1996), Kindle et al. (1989),

Cresswell et al. (1993), Hirst & Godfrey (1993), Molcard et al. (1994, 1996, 2001),

Miyama et al. (1995), Gordon et al. (1999), Potemra et al. (1997), Shriver &

Hurlburt (1997), Gordon & McClean (1999), and Potemra (1999).

Inverse solutions also show ITF transport toward the Indian Ocean: Piola &

Gordon (1984) 14 Sv; Toole & Warren (1993) 6.7 Sv; MacDonald & Wunsch (1996)

7 Sv; de las Heras & Schlitzer (1999) 13.2Sv; Ganachaud (1999) 15 ± 3Sv.

There is an urgent need for simulations of the Indonesian Throughflow (ITF)

using high resolution models. Though the limitations of regional models are well

known, they have their uses. This high resolution regional model properly resolves

all known important bathymetric features in the model domain such as sills, passages

and narrow channels and straits. As a result, important questions such as partition of

flow between passages, interaction of different water masses, influence of local winds,

turbulent mixing, and overall momentum and energy balances can be addressed.

Previous models have not had the resolution to properly account for the effect

of topography in the Indonesian Seas. Authors are in agreement that models with

higher resolution would improve results by including the effect of bottom topography.

Such models include the studies of Gordon & McClean (1999); Potemra et al. (2003);

Schneider & Barnett (1997); Schneider (1998); Verschell et al. (1995); Murray et al.

(1990); Kindle et al. (1989); Stammer et al. (2003); Schiller et al. (1998); Shriver

& Hurlburt (1997); Hirst fc Godfrey (1993); and Murtugudde et al. (1998) and are

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5

described in more detail in Chapter 2.

1.2. Objectives

The primary objectives of this research are to study:

• physical mechanisms controlling the splitting of the ITF between passages and the

effect of topography on flow paths taken;

• the effects of seasonal variation of transports through the open ports;

■ the influence of local winds on the regional circulation;

• the role of topography in the formation of distribution of 6 and S in the region with

an emphasis on the interaction of NPW and SPW;

• features of mixing and turbulence having an important effect on 6, S, and velocity

distributions.

1.3. A Summary of the Study

The study is divided into 3 parts. In the first part, Chapter 2, a description of

the model and properties of the simulated velocity fields are presented and discussed.

In the second part, Chapter 3, the influence of bottom topography on temperature

and salinity distributions is presented and discussed. In the third part, Chapter 4,

turbulence characteristics based on model results are presented and discussed.

In Chapter 2 the basic properties of a developed regional model of the circulation

of the Indonesian Seas are outlined. The model, based on the Princeton Ocean

Model, has 250 x 250 grid cells in the horizontal with grid spacing of ~ 10 km and

29 er-levels in the vertical. It is well known that the complex topography of the region strongly influences temperature, salinity and current distributions there. One

of the significant properties of this model is that all basic topographic features are

resolved. The model has four open ports to simulate inflow of North Pacific Water

from the Mindanao Current, inflow of South Pacific Water from the New Guinea

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 6

Coastal Current, outflow to the Pacific Ocean due to the North Equatorial Counter

Current, and outflow to the Indian Ocean due to the Indonesian Throughflow. Total

transports through the open ports and port normal velocities are specified from

observations. Orlanski’s conditions are employed at the open ports with port normal

velocity nudged to observed values and temperature and salinity to climatology. Port

channels are introduced so the effects of open boundary conditions do not impact

the dynamics of the main region. An additional friction was included in the vicinity

of some narrow passages and sills as a proxy for specific processes such as tides and

internal waves that occur within these topographic features.

Four experiments are discussed: seasonally varying and annual mean transports

and port normal velocities both with and without local winds. All experiments are

totally spun up after 10 years. This analysis uses data from the post spin up period

only. The basic properties of simulated total transports through the main passages

in the region, surface circulation and sea-surface heights are discussed. The portion

of North Pacific Water entering the Indonesian Seas relative to that leaving through

the North Equatorial Counter Current port is fairly constant throughout the year.

Most of this water takes the western route through the Makassar Strait. The portion

of South Pacific Water entering the Halmahera Sea compared to that exiting in the

North Equatorial Counter Current varies considerably with the seasons. Turning off

the local winds does not substantially influence the transport through main passages

in the model domain. Surface circulation patterns change substantially with the

seasons. The role of different terms in the heat and salt equations was investigated

by dividing the region into a number of boxes. For any given box, the sum of the

horizontal advective fluxes of temperature (salinity) through all sides of the box

is on the same order as the vertical heat (salt) flux at the surface, interior nudging

term, and the rate of time variation of the box integrated temperature (salinity). The

comparison of the basic structure of the model surface circulation, sea-surface heights

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7

and total transport values through the main passages with observations appears

satisfactory.

The main objective of the second part, Chapter 3, is to analyze the basic features

of potential temperature and salinity distributions in the Indonesian Seas simulated

by the model. The influence of bottom topography on the formation of temperature

and salinity distributions is considered by following the three major routes of flow

of North Pacific and South Pacific Water through the Indonesian Seas. Major eleva­

tions of bottom relief, such as the Sangihe Ridge; the topographic rise between the

Sulawesi Sea and Makassar Strait; the Dewakang Sill; the ridge between the Flores

and Banda Seas; the topographic rise between the Pacific Ocean and Morotai Basin;

the Lifamatola Sill; and the northern and southern Halmahera Sills; break the region

down to separate basins having different temperature and salinity stratifications.

The differences in stratification are caused by these topographic features that act

to impede the advection of cold and salty water from a basin (located upstream)

to the neighboring basin (located downstream). Arguments are included to support

this conclusion. In the upper ocean (500m), the Indonesian Throughflow is primarily

shaped by North Pacific Water taking the route through the Makassar Strait. Deep

Banda Sea water is formed by the overflow of North Pacific Water across the Lifam­

atola Passage into the Banda Sea. Below 500m South Pacific Water is blocked by

the Halmahera Sills and does not enter the Indonesian Seas. But in the upper ocean

(0-500m) SPW can probably penetrate into the Seram and Maluku Seas to mix with

NPW there.

There are no substantial structural changes of potential temperature and salinity

distributions between seasons, though values of some parameters of temperature

and salinity distributions (e.g., magnitudes of maxima and minima) can change.

It is shown that the main structure of the observed distributions of temperature

and salinity is satisfactorily displayed throughout the entire model domain. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 8

calculated transports of internal energy (heat) and salt mass through the Lombok

and Ombai Straits, and Timor Passage in August and February are in reasonable

agreement with published observed and simulated data.

In the last part, Chapter 4, aspects of turbulence and mixing in the Indonesian

Seas are presented and discussed. The results are based on the Mellor-Yamada 2.5

turbulence parameterization model. Though the importance of mixing in the In­

donesian Seas has been widely acknowledged, very few observations are available

and there have been no model studies of mixing or turbulent diffusion in the region.

The study is focused on turbulent diffusion and turbulent kinetic energy in the upper

mixed layer, the thermocline and in deep water near topographic features. Very large

turbulent kinetic energies and vertical turbulent diffusivities are seen around topog­

raphy and are important for the deep overflows found in the region. Large turbulent

energies and diffusivities found in the thermocline are important for the diffusion of

temperature and salinity signatures found in the Indonesian Seas. Monsoon winds

and local currents lead to large diffusivities in the upper mixed layer.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 9

CHAPTER 2

MODEL DESCRIPTION AND PROPERTIES OF THE

SIMULATED VELOCITY FIELDS

2.1. Introduction

This chapter outlines the main properties of a developed high-resolution baro-

clinic regional model of the Indonesian Seas circulation. This study is a continuation

of previous efforts summarized in Burnett et al. (2003) and Kamenkovich et al.

(2003). The model is based on the Princeton Ocean Model (POM). There are 29

cr-levels in the vertical and the horizontal resolution is ~ 10 km. Four open ports

are introduced to simulate the main currents entering and exiting the region of in­

terest. The primary motivation for developing such a model is to take into account

all basic features of the bottom topography in the Indonesian Seas region. Many

strong arguments demonstrate the importance of incorporating the relevant details

of the complex bottom topography to satisfactorily model temperature and salinity

distributions and currents in the region (see, e.g., Gordon & McClean 1999; Gordon

et al. 2003 and references given therein). The region is essentially a series of basins

separated by ridges with many sills,, passages and constrictions. To a great extent pe­

culiarities such as the sill at the northern end of the Sangihe Ridge, Labani Channel,

Dewakang Sill, Lombok Strait, North and South Halmahera Sills, and Lifamatola Sill

shape the partition of the main flow between different passages and the structure of

temperature and salinity in separate basins. But the width of these features ranges

from 30 to 50km so a rather fine resolution is required. We will show that with a

horizontal resolution of 10km we are able to properly incorporate all the essential

features of the bottom topography into the model.

A review of published modeling results for the Indonesian Seas show that most ef­

forts have been directed toward estimating total transport of the Indonesian Through-

flow (ITF) and its influence on the Pacific and Indian Ocean dynamics. Substantially

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 10

less attention has been paid to the analysis of the dynamics of the Indonesian Seas

per se, although the necessity of fine horizontal resolution has been noted in many

articles. We now summarize previous modeling studies of the ITF by briefly dis­

cussing baroclinic models with bottom topography and providing a brief summary

of other models. Kindle et al. (1989) investigated the mean, seasonal and interannual

variability of the ITF and forcing mechanisms responsible for this variability with

the Naval Ocean Research and Development Activity (NORDA) global model forced

by Hellerman & Rosenstein (1983) wind stress climatology and European Center for

Medium Weather Forecasting (ECMWF) 1000m6 winds for the period 1980-1987.

The model grid resolution was 0.5° in latitude and 0.7° in longitude. The NORDA

global model has multiple layers with the topography confined to the lowest layer.

The simulations yielded higher than expected transports through Lombok Strait be­

cause of coarse model resolution though total transports and current pathways are

very reasonable. The authors acknowledged that models with finer horizontal resolu­

tion should be utilized to investigate roles of the New Guinea Coastal Undercurrent

(NGCUC), intermediate and deep ITF and the effects of bottom topography on the

circulation. Hirst & Godfrey (1993) examined the role of the ITF in an Ocean Global

Circulation Model (OGCM), the main aim of which was to understand the role of the

throughflow in the mean mass and heat balances of the Indian and Pacific Oceans.

They used the Cox version of the Bryan-Cox OGCM. The horizontal grid spacing

was 1.5928° in latitude and 2.8125° in longitude. Schneider & Barnett (1997) were

the first to examine the ITF with a global coupled ocean-atmosphere model. The

horizontal resolution of the ocean model was 0.5° within 10° of the equator increasing

to 2.8° poleward of 20°. The model had two entrance paths to the Indonesian Seas,

one of which is the Torres Strait, and two exit passages. Schneider (1998) modeled

the role of the ITF in the global climate system using the same model. The au­

thor acknowledges that the coarse resolution of the coupled model does not resolve

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 11

the complicated topography and pathways in the Indonesian Seas and the problem

needs to be addressed by running simulations with higher resolution. Shriver &

Hurlburt (1997) used a set of global ocean simulations to investigate: (i) pathways

feeding the ITF at different depths and (ii) the role of the thermohaline circulation

there. The model had horizontal resolution of 0.5° with vertical resolution rang­

ing from 1.5-layer reduced gravity to six layers. It is possible that confining rough

topography to the bottom layer reduces the effect of bottom topography. Gordon &

McClean (1999) used the output of the Los Alamos Parallel Ocean Program (POP)

1/6° global model forced by ECMWF wind stresses for the period 1985 to 1995 to

compare the simulated results for the Indonesian Seas region with observations made

during the southeast monsoon (August 1993) and the northwest monsoon (Febru­

ary 1994). They found some major discrepancies between observations and model

results for potential temperature and salinity distributions largely attributable to

model resolution, e.g., the Dewakang Sill (depth ~ 550m) was replaced by a sill of

~ 200m depth and the Torres Strait is too deep and too wide. They suggest that

some model inadequacies will be improved by increasing model resolution. Schiller et

al. (1998) used the Geophysical Fluid Dynamics Laboratory Modular Ocean Model2

OGCM (GFDL MOM2 OGCM) to model near surface dynamics and thermodynam­

ics in the Indian Ocean and the ITF. The model resolution is 2° in longitude and

0.5° in latitude within 8° of the equator increasing to 5.85° at the poles. They in­

cluded increased viscosity in the region, centered in the Banda Sea and decreasing in

value outward from there, to account for large tidal currents found over the irregular

topography there. Though their estimates of transport are not unreasonable, they

have a reduced number of paths through the Indonesian Seas. Potemra et al. (2003)

examined the exchange of water between the Pacific and Indian Oceans via the ITF

and its temporal variability using output from the Parallel Ocean Climate Model

(POCM). The model is a global-scale ocean model having horizontal resolution of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 12

~ 0.4° and 20 z-levels. However, it did not resolve flow through the Lombok Strait

and consistently simulates higher salinities, possibly due to too much salt entering

the region through the Torres Strait. Their Figure 1 also shows the consequences of

limited model resolution at the entrance and exit straits to the region. Stammer et

al. (2003) used the results of a fully constrained GCM to estimate ocean transports

of volume, heat and freshwater and their temporal changes including estimates of

the ITF transport. All estimates are based on the Massachusetts Institute of Tech­

nology OGCM (MIT OGCM) and its adjoint (see Marshall et al. 1997 and Marotzke

et al. 1999). The model has horizontal resolution of 2° and 22 vertical levels. The

authors acknowledge that flow estimates are limited by the low resolution of the

model. The following authors have successfully applied the reduced gravity model

with no motion in the lower layer to the analysis of the Indonesian Seas circulation.

Kindle et al. (1987); Godfrey (1989; 1996); Murray et al. (1990); Inoue & Welsh

(1993); Verschell et al. (1995); Qiu et al. (1999); and Murtrrgudde et al. (1998).

The bottom topography was essentially ignored in these simulations though Godfrey

(1996) and Godfrey & Masumoto (1999) did try to estimate its effect.

To achieve a more detailed understanding of the Indonesian Seas dynamics we

have attempted to separate the global and regional problems. It is well known that

regional models have limitations but we are confident that questions such as the par­

tition of the main flow between different passages, the influence of local winds, the

interaction of North and South Pacific Waters, overall momentum and energy bal­

ances, and properties of turbulent mixing in the region can be successfully addressed

with this model.

In Section 2 the model configuration is outlined. In Section 3 some of the model

results (total transports through main passages, surface circulation, sea-surface heights

(SSH)) are discussed. Estimates of terms in the heat and salt diffusion equations

are presented in Section 4. Section 5 focuses on the comparison of simulated and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 13

observed data. Finally, Section 6 summarizes the basic results of this study.

2.2 Model Description

The horizontal model configuration coincides with that described in Burnett et

al. (2003; Figures 1 and 2). The model domain and bathymetry are shown in Figure

2.1(a) and the names of important topographic features and open ports are given in

Figure 2.1(b). The domain extends throughout the Indonesian Seas region and has

250x 250 grid cells in the horizontal, identified in Figure 2.1(b) by i and j. These grid

cells are ~ 10km long in both horizontal directions and have been generated using the

GRID program provided with the POM (see Mellor 2004). The model contains four

open ports to simulate the impact of major currents entering and exiting the region.

The orthogonal curvilinear coordinate system used in the POM has been rotated

relative to latitude and longitude so that the open port in the Indian Ocean region

(10 port) coincides with a transect line extending from Java to northwest Australia

along which observations were made (see, e.g., Fieux et al. (1994), Sprintall et al.

(2000)). The ITF exits the model domain and enters the Indian Ocean through this

port. There are three open ports in the Pacific region: the Mindanao Current port

in the north (MC port); the New Guinea Coastal Current port (NGCC port) just

to the north of New Guinea/Irian Jaya; and the North Equatorial Counter Current

port (NECC port) in the east. Pacific waters enter the domain through the MC and

NGCC ports and exit through the NECC port. The grey regions in Figure 2.1(a)

and 2.1(b) are shallow (depth is less than 100m) and are considered as land (see

discussion of model topography below). There are 29 a-levels in the vertical so that

important features of the vertical structure are properly resolved over all types of

topography.

2.2.1. Bottom Topography

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 14

Consider two characteristics of the variability of depth H(i,j): slope in the

x-direction, sx(i,j), and in the y-direction, sy(i,j), which are defined as

The standard recommendation by Mellor et al. (1994) is that the maximum

value of sx and sy should be less than 0.4 over the entire model domain. Otherwise,

the model will not work properly because of the pressure gradient problem. We

started with the ETOP05 dataset, which we interpolated onto the model grid and

calculated maxima of sx and sy over all grid points with > 0.1m. The

maximum value of sx was 1.74 (77(144, 225) = 133m; 77( 143,225) = 1923m) while

that of sy was 1.80 (/T(102, 96) = 2606m; 77(102, 95) = 135m) and the maximum

depth was 77( 165,226) = 9302m. So, to follow the recommendation we needed to

modify the bottom topography.

First, the lOOm-isobath was chosen as the coastline. This choice closed the pas­

sages to the Sulu Sea, , Torres Strait, and passages around north Irian Jaya.

Small passages in the Lesser Sunda Island Chain were closed but the important Lom­

bok and Ombai Straits were retained. Then, without changing the structure of the

topography, we introduced some local modifications in 77 to eliminate overly large

local slopes. For example, we set the maximum depth in the model domain to 6000m

(e.g., the Mindanao Trench is deeper than 9000m); set minimum depth of seamounts

in the Pacific basin to 2000m; removed two islands west of the NGCC port, and so

on.

Second, we introduced a rigid boundary around the model domain of 5-cell width

except at the four open port channels (Figure 2.1); eliminated depth variation along

the port channels; and removed all 1-point, islands.

Third, we smoothed this modified bottom topography by applying the following

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.1. Model bathymetry and domain (a) (above) and names of important topographic features, open ports and islands (b). Note that in (b) the model domain is shown in x, y coordinates (i, j are given), and the depth in the grey region is less than 100m. Notice also that the island chain extending from Java in the west to Timor Island in the east is the Lesser Sunda Island Chain.

formula (Il’in et al. 1974),

(2 .2 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 16

250

MindanaoCurrentport

Sulawesi Pacific Sea Ocean 200 Morotai North Equatorial BasML* Countercurrent , ^^®Waluku SeoJKt port M il' Lifamatola m ^Passag^ New Guinea Coastal 150 Current port \ Taliabu y/^Seram Sea Mangole Buru s 3

Banda Sea 100 Indian Ocean port Timor Passage

50

Figure 2.1(b).

where H is the smoothed depth, P is the grid point, a and R are some constants,

Pm is the point whose distance rm from P is less than /?. and M is the total number

of such points. The distances R and rm are scaled by the Earth’s radius; the rm are

calculated after mapping the domain on the Mercator plane.

We used a — 2R2. The first smoothing was made with R = 0.02 (a smoothing

radius of ~ 128 km). We then restored the unsmoothed bathymetry in all pas­

sages/sills/straits that, had been oversmoothed with R = 0.02. The next smoothing

was done with R = 0.005 (a smoothing radius of ~ 32km). We then again restored

the unsmoothed bathymetry for all passages/sills/straits that had been oversmoothed

with R = 0.005 and made another smoothing with R — 0.0025 (a smoothing radius of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 17

~ 16km). Finally, we applied the POM GRID subroutine SLPMIN twice to further

reduce large slopes (slmin —0.19).

This smoothed topography was used in all numerical experiments. For this topog­

raphy the maximum value of sx was equal to 0.41 (if (141, 207) = 638m; H (140, 270)

= 970m) while the maximum value of sy was equal to 0.40 (if(8 ,116) = 332m;

if (8,115) = 501m). The number of ocean cells is 31105 (approximately 50% of

all grid points). The maximum depth is Hmax(209, 245) = 5999m. Does such a

smoothed bottom topography retain the dynamically important features? From our

knowledge of the general pattern of the circulation in the Indonesian Seas we conclude

that the topography in the areas of the Sangihe Ridge, Labani Channel, Dewakang

Sill, Halmahera Sea, Lifamatola Passage, Lombok Strait, Ombai Strait, Timor Pas­

sage, and the passage between Irian Jaya and Seram Island is critically important.

It is shown in Appendix A that the basic features of the bottom topography in these

subregions are retained.

2.2.2. Boundary Conditions

We describe here the open boundary conditions only. The remaining boundary con­

ditions are standard for the POM and are given in Appendix B for the sake of

completeness.

2.2.2.1. Boundary conditions for the depth-averaged motion.

1. Specification of the normal velocity based on the prescribed total transport and

assumed simple distributions of this velocity across the port.

2. Tangential velocity is equal to zero.

2.2.2.2. Boundary conditions for the 3D motion.

1. Orlanski’s condition with nudging for the normal velocity. Following the recom­

mendations of Marchesiello et al. (2001), we have for the Indian Ocean IO port:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 18

«(n+1) (2, j, k) — u{n ^ (2, j, k) 2At = ,. m uM(3, j,fc) ~ 0.5[«(n+1)(2,j, fc) + u ^ 1)(2,j,fc)] __

'•fo[“‘”_1)(2, J, *) - *)]. (2-3)

where c(j\ fc) is calculated according to the P0M2K code (see Orlanski 1976); At is

the time step; AXj is the distance between (2 ,j) and (3 ,j) grid points; superscript n

indicates the time nAt; rjQ is the nudging coefhcient for the normal velocity u; ue is

the port normal velocity prescribed from observations. This formula is used for the

calculation of the boundary value rdn+1)(2, j, k). The nudging coefficient differs

for inflow and outflow points, identified by the sign of ffin)(2, j, k)—c(j, k): for outflow

points (2, j, k) — c(j, k) < 0 while for inflow points u ^ (2,j,k) — c(j, k) > 0. Note

that by definition c(j, k) > 0. Usually the nudging coefficient for inflow points is

larger than for outflow points and to provide a smooth transition from inflow to

outflow points we used the following formula

r^o = rout + 0.5rin(l + tanh[a(u(n)(2, j, k) - c(j, k) - ue)]), (2.4)

where rottt = 10~4s_1; rin = 5 • 10-4s -1; a = 60sm_1; ue = 0.05ras_1. Similar

formulas have been used for calculating ffin+1)(im, j, k) for the NECC and NGCC

ports and v^n+1^ (i, jm, k) for the MC port. Note that im = 250, jm = 250 and k

indicates a*, level.

2. Tangential velocity is set to zero.

3. Orlanski’s conditions with nudging for temperature and salinity. Basically these

conditons are similar to (2.3), where, for example, T (l, j, k) and k) are nudged

to the corresponding climatological values, and the nudging coefficient r \SQ is given by

the same formula, (2.4), but with rout = 5- 10_8s_1; rin = 10-5s-1. Similar formulas

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 19

have been used at the NGCC, NECC, and MC ports.

4. The turbulence energy and length scale are set to zero.

2.2.3. Modifications of the POM Basic Equations

2.2.3.1. Within Port Channels. The port channels have been introduced to allow

for tapering off of any possible effects of nudging at the boundary. Therefore the

basic equations have been modified within the port channels in the following way.

1. On the right-hand side of the momentum equations we added frictional, terms

—r\v\u, v) for the outflow port channels (10, i = 3 — 5; and NECC, i — 246 — 249),

nudging and frictional terms —r^(u — ue, v ) for the inflow port channel NGCC (i —

246 — 249) and —rjV\(u, v — ve) for the inflow port channel MC (j = 246 — 249), with

decreasing toward the main region.

2. On the right-hand side of the temperature and salinity equations we added nudging

terms —rls\ T —Tdim); —r\s\S — Sdim) for the IO port channel (i = 2 — 5), for NGCC

port channel (i — 246 — 249), and so on, with decreasing toward the main region.

2.2.3.2. In the Mam Domain.

1. To control the partition of flow in some important regions we incorporated ad­

ditional frictional terms of the form —r^(x,y)u,—r^(x,y)v into the momentum

equations, with fiv\x,y) non-zero in the vicinity of some passages and sills. Details

are given in Table 1, This additional friction is a proxy for some specific processes

(tides, internal waves) that occur within the narrow passages and at sills. The

approach is rather crude but these processes somehow need to be taken into ac­

count. Otherwise, due to the western intensification, we obtained overly strong flows

through, for example, the Lombok and Ornbai Straits. A similar approach was taken

by Schiller et al. (1998).

2. On the right-hand side of the temperature and salinity equations we introduced

terms - r (0)(T - Tclim); - Sdim), where Tdim and Sdirn are the corresponding

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 20

Location subregion maximum value (s Lombok i=5-22; j=109-125 5 x 10“5 southwest Makassar i=29-43; j=135-149 5 x 1(T5 Ombai i=90-102; j=86-94 2 x i t r 5 Guinea-Seram passage i= 160-167; j=129-140 2 x i t r 4 Taliabu-Sulawesi passage 1=102-118; j=151-167 2 x i t r 4 Obi-Halmahera passage i=139-155; j=146-162 itr5 Lifamatola passage i= 130-144; j=144-162 i t r 4 southwest Flores Sea i=44-50; j=108-125 i t r 4 Buru-Seram passage 1=131-147; j = 128-138 5 x i t r 5 Obi-Seram passage 1=142-152; j=134-151 2 x i t r 5

Table 1. Subregions of the model domain where additional horizontal friction has been applied as follows: —r(i,j)u(i,j,k) and — r(i, j)v(i, j, k). The values of r are maximal at the centers of the subregions and decay exponentially toward the bound­ aries.

climatological values. The coefficient r ^ is a function of a: it is equal to 10-3s-1 at

the surface and then decays exponentially to constant value of l(T"8s_1 at cr < —0.04

(e.g., it will be 10-8s-1 for depths greater than 40m if the total depth is 1000m).

These terms introduce additional heat and fresh water fluxes basically within the

upper layer, attracting temperature and salinity to seasonally varying climatological

values. They were required for two reasons. First, the surface fluxes of heat, and

especially fresh water, are highly variable and poorly known from observations. Sec­

ond, utilization of climatological winds in the model results in underestimation of

vertical mixing and depth of the mixed layer by the Mellor-Yamada parameteriza­

tion. Moreover, it is very difficult to model a very complicated structure of salinity

in the upper layer. Nudging of the upper layer temperature and salinity toward

climatological data appeared to be a more reliable way to set up surface boundary

conditions for temperature and salinity. Note that a similar expression was used by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 21

Haney (1971) in formulating the boundary condition for the heat equation at the

free surface.

3. To approximate temperature and salinity advection we used the Smolarkiewicz

scheme with two iterations and smoothing parameter sw = 0.5. Some modifications

were introduced in the calculation of the buoyancy frequency following basically

Hunter’s suggestion (OZPOM code). All other parameters were chosen in accordance

with the standard version of the POM2K code. Note that the horizontal diffusivity

is less than the horizontal friction coefficient by a factor of 5.

2.2.4■ The Specification of the Port Normal Velocity ue and Total Port Transports

The grid spacing of the Levitus climatology data does not allow for use of

geostrophic relations to specify ue. Moreover, the NGCC port is located too close

to the equator. Therefore, published observations and some modeling results have

been used to determine the port normal velocity ue and total port transports. We

will estimate annual mean and August and February values of port transports and

port normal velocity distributions. Time deviations from the annual mean of port

transports and port normal velocity were approximated by the cosine of the annual

period.

2.2.4-1- 10 port. The open 10 (Indian Ocean) port is basically the region to which

the net westward transport is confined. First, we will model the mid-port normal

velocity profile um(z ) using the geostrophic calculations of Fieux et al. (1994; see

their Figs 12 and 13) based on the JADE (Java Australia Dynamic Experiment)

data. This profile is shown in Figure 2.2(a). The velocity is everywhere directed

toward the Indian Ocean. A maximum magnitude of 10 cm/s is seen at the surface

decreasing monotonically to zero below 450m. The profile um(z) does not change

over the course of a year. Second, to calculate ue(y,zfi) we apply the following

formula

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 22

V 0 500 1000 1000 - 2000 — annual mean 1500 — August S 3000 2000 ■o — - February 4000 2500 3000 5000 (a) 3500 -0.1 -0.05 0

0

1000 1000

§2000 2000 £ 3000 .§ 3000 4000 4000 5000 5000 0.2 ■0.5 -0.4 -0.3 -0.2 -0.1 0 velocity (m/s) velocity (m/s)

Figure 2.2. Modeled mid-port normal velocity profiles at the IO port (a), NGCC port (b), NECC port (c), and MC port (d). The profiles at (a), (c) and (d) are assumed time independent. Profiles for August and February are also given for (b).

Ue(y,z,t) = 0 m " Sx(y,t), (2.5) J~H Um(z)az

where Sx(y,t) is the depth-integrated port normal velocity determined by using the

specified total transport and assumed simple distribution of velocity across the port.

We have chosen an annual mean throughflow transport of 12. 5Sv with a maximum

of 17.05T in August and a minimum of 8 .OSv in February. Observations of the

transport of the Indonesian Throughflow have been made by several investigators.

Fieux et al. (1994) observed a transport 18.6±7.0ST across the Bali-Australia section

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 23

in August 1989. The transport of the ITF consists of the transports through the

Lombok and Ombai Straits and the Timor Passage. Molcard et al. (1996) estimated

a mean transport through the Timor Passage of 4.3 ± LOST in the upper 1250 db

and 2.5 ± 0.8 Sv in the upper 500db with a maximum of 3.0 — 4.05T in August and

a minimum of 1.0 — 2.0ST in February. The measurements were made from current

meter moorings between March 1992 and April 1993. Molcard et al. (2001) estimated

a mean transport through the Ombai Strait of 5.0 ± 1.QST with a maximum of

7.0 —8.0ST in August and a minimum of 4.05T in February. The measurements were

made from a current meter mooring deployed for one year in December 1995. Murray

& Arief (1990) made both observations and model calculations of the transport

through the Lombok Strait. They observed a mean annual transport through the

Lombok Strait of 1.8ST with a maximum of LOST in August and a minimum of

LOSe in February. The observations were made from current meter arrays deployed

in the north Lombok Strait from January 1985 to March 1986. Model results of

Murray & Arief (1990) gave a mean transport of 2.0ST with a maximum of 3.5 Sv

in August and a minimum of LOST in February. Masumoto & Yamagata (1996)

modeled the seasonal variability of the transport of the ITF and found an annual

mean transport of 9.5 Sv with a maximum of 11.6Se in August and a minimum of

6.0Se in February.

2.2.4-2- NGCC port. We used the data shown in Figures 3, 8 and 9 of Ueki et al.

(2003) to model the typical mid-port normal velocity profile um(z,t ) at the entrance

to the NGCC (New Guinea Coastal Current) port. Profiles for August, February

and the annual mean are shown in Figure 2.2(b). The velocity profile changes with

the seasons. The observed near surface velocity of the NGCC is equal to 0.3 rn/s and

directed westward in summer, is equal to 0.1 m/s and directed eastward in winter, and

is equal to 0.1 m/s and directed westward in the annual mean (all estimates here and

in what follows are approximate). The NGCUC (New Guinea Coastal UnderCurrent)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 24

core (200m) remains at the same position throughout the year with a maximum

velocity of 0.6m/s directed westward in summer, OAm/s directed westward in winter,

and 0.5 m/s in the same direction in the annual mean. Magnitude decreases to zero

below 400m. The current is about 100 km wide throughout the year which agrees with

the approximate width of our port. These values are in good agreement with other

available observations. To determine port normal velocity ue(y,z,t ) a procedure

similar to that described for the 10 port was applied.

Ueki et al. (2003) deployed a series of ADCP moorings off the coast of New Guinea

for 5 years to investigate the seasonal and interannual variations of the NGCC and

the NGCUC. Two of the ADCP moorings were located off the New Guinea coast

at 2°S', 142°E- and 2.5°S 142°JF, 400 km to the east of our open NGCC port. We

assumed that these measurements provide the correct structure and transport of

currents entering our domain through the NGCC port. The system of currents

outside this port is extremely complicated. The South Equatorial Current (SEC)

splits into several branches as it approaches the southeastern edge of New Guinea.

Part of the SEC is returned eastward through the Solomon Strait as a branch of the

South Equatorial Counter Current (SECC) while the remainder continues westward

through either the Vitiaz Strait or St. George’s Channel (see, e.g., Lindstrom et

al. 1990). Further west another portion of the SEC peels off and returns eastward

around Manus Island forming the New Ireland Coastal UnderCurrent (NICU) that

interacts with the Equatorial Undercurrent (EUC) as it approaches the equator (see,

e.g., Butt & Lindstrom 1994). It seems that there is no further peeling off of the

SEC between the NICU and the location of our port (see, e.g., the schematics of

Lukas et al. (1996) and Godfrey (1996) (Figure 1 in both cases), and papers by

Gouriou & Toole (1993), Qiu & Joyce (1992), and Toole et al. (1988) for a further

discussion). Figure 3 of Ueki et al. (2003) shows the structure of the annual mean

NGCUC and NGCC at the two ADCP moorings located at 142° E. Ueki et al. (2003)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 25

also conducted 11 shipboard ADCP and CTD cruises in summer and winter between

1994 and 2000 to investigate the seasonal variability of the structure of these currents

(see their Figs 8 and 9).

We have chosen an annual mean volume transport through the NGCC port of

22.5Sv, a summer transport of 28.0£u and a winter transport of 17.0ST. We made

seasonal and annual mean estimates of volume transport at the NGCC port, respec­

tively, from the Ueki et al. (2003) calculations based on the summer and winter

shipboard cruise data and the ADCP mooring (see their Figs 12 and 17). Several

other investigators have made observations of the NGCC and NGCUC. Lindstrom

et al. (1987) made several transects of the NGCC and NGCUC during the Western

Equatorial Pacific Ocean Circulation Study (WEPOCS), where the seasonal variabil­

ity of Low Latitude Western Boundary Currents (LLWBC) was investigated. Two

expeditions were conducted during June-August 1985 (WEPOCS I) and January-

February 1986 (WEPOCS II) where special attention was paid to sampling of the

LLWBC along the coast of Papua New Guinea. In May 1988, Butt & Lindstrom

(1994) made ADCP and CTD measurements off the coast of New Guinea to investi­

gate the structure of the SEC and EUC system there. Murray et al. (1995) provided

annual transport measurements of the SEC at the Vitiaz Strait from five current

moorings deployed from February 1992 to April 1993. We have chosen to use the

observations of Ueki et al. (2003) since they represent the most recently updated

dataset in this region.

2.2.4-3. NECC port. We have modeled the mid-port normal velocity um(z) at the

NECC (North Equatorial Counter Current) port using the geostrophic calculations

of Qiu & Joyce (1992). The profile is shown in Figure 2.2(c). It has the same

structure throughout the year. The velocity is toward the west at all depths with

a maximum value of 27 cm/s in the upper 70m in the annual mean. This value

decreases monotonically to 5cm/s at 350m and drops off to zero at 800m. This

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 26

structure and magnitude is in good agreement with other available observations. To

determine port normal velocity ue(y,z,t ) a procedure similar to that described for

the 10 port was applied.

Qiu & Joyce (1992) analyzed 22 years of data from hydrographic cruises col­

lected by the Japanese Meteorological Agency along 137°E in winter (1967-1988)

and summer (1972-1988) to study the interannual variability of the western North

Pacific. Note that the 137° E meridian is located adjacent to the NECC port (Qiu

& Joyce 1992; Figure 1). The hydrographic sections were made from 34°A to 1°S

thereby crossing the full extent of the NECC. Hydrographic stations were spaced

approximately 0.5° apart in the region of the NECC. Qiu & Joyce (1992) results of

the geostrophic calculations relative to lOOOdb between 2 °N and 34°^ are shown in

their Figure 4. They found that the mean position of the NECC is between 2° N

and 7 °N. The mean cross-section of the NECC shows a velocity core of more than

0.5m/s centered at 50m and located between 4 °N and 5 °N (see Figure 4 of Qiu

& Joyce (1992)). The current is symmetrical about the core down to 150m, be­

low which it is offset to the south. The geostrophic current extends to a depth of

about 750m but is less than 0.2 m/s below 250m and less than 0.1m/s below 500m.

Other geostrophic calculations include those of Gouriou & Toole (1993) who used a

similar data set to that of Qiu & Joyce (1992). Gouriou & Toole (1993) analyzed

hydrographic data collected by the Japanese Meteorological Agency between 1972

and 1985 along 137°E, using reference levels of 600d6 and 1000 db. They found the

mean position and structure of the NECC to be similar to that found by Qiu & Joyce

(1992) (see Figure 4c of Gouriou & Toole 1993).

We have chosen an annual mean transport of JOS'?;, 46ST in summer and 35 Sv

in winter. Qiu & Joyce (1992) calculated a mean eastward geostrophic transport of

42.1ST. Gouriou & Toole (1993) calculated annual mean NECC geostrophic trans­

port of 43.1 and 48.2ST, summer transport of 52.5 and 56.3ST, and winter transport

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 27

of 34.9 and 44.8 Sv referenced to 600 and lOOOdb, respectively.

2.2.4-4- MC port. Because of the low annual variability of the Mindanao Cur­

rent (MC) we used a combination of the ADCP observations and the geostrophic

calculations of Lukas et al. (1991) at 10° N (see their Figures 6 and 7) to model

the mid-port normal velocity profile um(z). The profile is shown in Figure 2.2(d).

It does not change throughout the year. The velocity is everywhere into the model

domain with magnitudes exceeding 50 cm/s in the upper 250m, decreasing to 20cm/s

at 500m and is zero below 900m. This structure is in good agreement with other

available observations. To determine port normal velocity ue(y,z,t ) a procedure

similar to that described for the IO port was applied.

Lukas et al. (1991) observed the MC during the WEPOCS in June and July 1988.

Data were collected from drifter launches, ADCP observations and CTD02 casts,

including zonal measurements across the MC at 12° N, 10°N, 8°N, 7°N, 6°N (see

Figure 1 of Lukas et al. 1991). The MC flows along the Philippine coast from about

13.5°N to the southern tip of Mindanao Island at 6°Ar, where part of it extends out

into the Sulawesi Sea (see Figure 5 of Lukas et al. 1991). The current is about lOOfcm

wide at 10° iV near the location of our open MC port. The current stays close to the

coast with velocities concentrated in the upper 200m and a maximum core located at

about 100m (see Figure 6 of Lukas et al. 1991). Geostrophic calculations show that

the current extends to a depth of about 700m: (see Figure 7 of Lukas et al. 1991). The

MC is mainly supplied by the North Equatorial Current (NEC) between 12°^ and

10°N (see Figure 4 of Lukas et al. 1991). Observations at different times of the year

show the MC to be very stable. Wijffels et al. (1995) investigated the mean structure

and variability of the MC at 8°N based on eight surveys between September 1987

and June 1990 and found that the MC is remarkably steady in position and structure

over that time period. Kashino et al. (2005) investigated the variability of the MC

from mooring observations between October 1999 and July 2002 and basically came

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 28

to the same conclusion.

We have chosen an annual mean transport of 30 Sv, 35Sv in August and 25 Sv

in February. Wijffels et al. (1995) estimated a mean transport of 25 ± 4 Sv in the

thermocline and above and 39 ± 105T above 2000db with a maximum in summer

and a minimum in winter. Qu et al. (1998) provided mean transport estimates of

27 — 28Sv from CTD measurements made every September-October between 1986

and 1990. Transport estimates from model results include those of Qiu & Lukas

(1996) who estimated a mean transport of 25 — 40ST increasing by 5 Sv in summer

and decreasing by 5Sv in winter with 1557; transported within the Mindanao eddy.

The model of Yaremchuk & Qu (2004) gave a mean transport of 30 Sv with 355c in

summer and 25 Sv in winter.

2.2.5. Period of integration

To monitor the evolution of experiments two characteristics Ey and Es appeared

useful,

Ey = X (2 .6)

where Ey is integrated over the entire volume of the domain and Es over the horizon­

tal area of the domain; pm is the mean density determined by averaging the annual

climatological density over the horizontal; N is the buoyancy frequency calculated

from pm; p' — p — pm; wpom is the POM vertical velocity (see, e.g., Appendix D); g

is gravity. Note that these characteristics are not exactly equal to the total energy

(kinetic and potential) and total potential energy created by the SSH variations.

Table 2 summarizes the integration periods, port transports and port normal

velocity profiles for the experiments that were performed. The time deviations from

the annual mean of temperature, salinity, and wind stress components were

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 29

Port transports Velocity Period Type of Sv (x 106m3/s) profiles of inte­ Ex solution MC NGCC NECC 10 at ports gration annual mean, annual mean 21 1 with local wind -30.0 -22.5 40.0 12.5 (see Figure 2) years annual mean, 2 no local wind the same as in exp 1 August mean, August mean 6 3 with local wind -35.0 -28.0 46.0 17.0 (see Figure 2) years August mean, 4 no local wind the same as in exp 3 February mean, February mean 6 5 with local wind -25.0 -17.0 34.0 8.0 (see Figure 2) years February mean, 6 no local wind the same as in exp 5 seasonal with Annual mean plus cosine Annual mean 12 7 local wind of annual period plus cosine of years based on annual period annual mean (see Figure 2) and 2 Fourier harmonics seasonal with 8 no local wind the same as in exp 7

Table 2. List, of experiments. Transports out of the model domain are positive; transports into the model are negative. In the first column 9- , S'-climatology, and wind stresses are characterized.

approximated by two Fourier harmonics. Here, we analyze experiments 1, 2, 7, and

8 (annual mean with and without wind and seasonally varying with and without

wind). Experiments 3, 4, 5, and 6 were used for technical purposes. Figure 2.3

shows variations of Ey and Eg with time for experiments 1 and 7. Experiments 7

and 8 were run for 12 years and experiments 1 and 2 for 21 years. Figure 2.3 clearly

shows that these experiments have been completely spun up after 10 years.

To estimate a possible effect of the POM approximation of horizontal pressure

gradients we ran, as is usually done, some experiments without any external factors

and with distribution of potential temperature and salinity dependent on z only. The

errors were not substantial.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 30

x 10

0 2 4 6 8 10 12 14 16 18 20 years

x 10 14

12

>> 0) 8 0 C UJ 6 v ^ •* ,/*v* * rs» * *"V‘sr

4, 0 2 4 6 8 10 12 14 16 18 20 years

Figure 2.3. The evolution of E v and Es with time. The characteristics Ev are shown in (a) for the seasonally varying experiment with local wind (exp 7; solid line) and the annual mean experiment with local wind (exp 1; broken line). The characteristics Es are shown in (b) for the same two experiments as (a). Variability of the integrated value of Ev in exp 1 is ~ 0.1 x 10 17kgrri2/s2 and for Es is ~ 0.5 x 1013%m 2/s 2.

2.3. Results

As seen in Figure 2.4(a), Pacific water enters the Indonesian Seas region through

the MC and NGCC ports. North Pacific Water (NPW) enters the region through

the MC port before splitting into three branches. The first branch of NPW follows a

western path, flowing over the Karakelong and Sangihe Ridges, through the Sulawesi

Sea, and along the Makassar Strait, at the southern part of which it diverges. Part

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 31

of this flow exits the region through the Lombok Strait, but the main portion passes

through the Flores Sea to the Banda Sea. The second branch of NPW takes the north-

south route through the Morotai Basin, Maluku Sea, Lifamatola Passage, Seram Sea

and finally into the Banda Sea, where it meets with water from the western route.

The third branch of NPW leaves the region through the NECC port. South Pacific

Water (SPW) enters the region through the NGCC port before splitting into two

branches. The first branch of SPW flows through the Halmahera Sea to the Seram

Sea where it meets with NPW taking the north-south route. The second branch of

SPW leaves the region through the NECC port. Finally, Banda Sea water exits the

region through the Ombai Strait and Timor Passage.

2.3.1. Transports

We now consider seasonal variations of the simulated transports in experiments

7 and 8 and compare them with those of experiments 1 and 2. First, we examine

transports through the inflow open ports and main passages for experiment 7 (sea­

sonally varying circulation with wind) for August and February and for experiment

1 (annual mean circulation with wind), shown in Figures 2.4(a)-(c), respectively.

Transports through the NIC (inflow of NPW) and NGCC (inflow of SPW) ports as

well as through the Makassar Strait, Maluku Sea, Halmahera Sea, Lombok Strait,

Ombai Strait, and Timor Passage are stronger in August and weaker in February,

when compared with the annual mean. Variations relative to annual mean are ±17%

and ±24% for the inflow through MC and NGCC ports, respectively. Correspond­

ingly, the increase (decrease) of outflow transports through the NECC and 10 ports

in August (February) is observed (±15% and ±36% of the annual mean values for

NECC and 10 ports, respectively). A portion of the MC transport flows into the

Indonesian Seas region while the remainder outflows through the NECC port. In

August the total transport of NPW entering the Indonesian Seas region through the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 32

50 100 150 200 250 j

Figure 2.4(a). Transports across several sections of the model domain including the four open ports. Values for experiment 7 (seasonally varying with local wind) for August and February, and experiment 1 (annual mean with local wind) are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 (seasonally varying with no local wind) for August and February, and experiment 2 (annual mean with no local wind) are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration. Some arrows are slanted for graphic clarity.

Halmahera-Mindanao section (i — 154), is 12.2Sv (35% of total transport of NPW

through the MC port), while in February the corresponding value is 7.6 Sv (30%

of total transport of NPW through the MC port). Of these values, 78% of NPW

is transported through the Makassar Strait in August and 68% in February. For

experiment 1 a transport of 9ASv enters the Indonesian Seas (31% of the total

transport of NPW through the MC port) with 7.0 Sv flowing through the Makassar

Strait (74% of the transport of NPW into the Indonesian Seas). Note that the

percentage of NPW entering the Indonesian Seas region does not change noticeably.

Unlike the transports of NPW, the percentage of SPW entering the Indonesian Seas

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 33

Figure 2.4(b).

varies substantially. In August 4.8ST of SPW enters the Halmahera (17% of total

transport of SPW through the NGCC port), while in February the corresponding

value is 0.45?; (2.4% of total transport of SPW through the NGCC port). The

annual mean value is 3.1 Sv (14% of total transport of SPW through the NGCC

port). Essentially all of the NPW transported through the Maluku Sea enters the

Seram Sea by way of the Lifamatola Passage. This flow combined with that through

the Halmahera Sea accounts for the total transport into the Banda Sea from the

north-south and eastern routes, amounting to QASv in August, 2.7 Sv in February,

and 4.9 Sv annually. Almost all of this water enters the Banda Sea through the

Mangole-Buru section (i = 129).

It is useful to compare transports through the Makassar Strait and the Mangole-

Buru section. The latter is 67% of the former in August, 52% in February, and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 34

Figure 2.4(c).

70% annually. The comparison of transports through the Lombok Strait and Flores

Sea section (i = 61) provides the following numbers. The transport through the Lom­

bok Strait accounts for 22%, 25%, and 23% of the transport through the Makassar

Strait in August, February, and annually, respectively, with the remainder entering

the Banda Sea through the Flores Sea section.

Relative to the annual mean, the increase (decrease) of transport through the

Lombok Strait, Ombai Strait, and Timor Passage are 31% (19%), 8 % (4%), and

62% (69%) in August (February), respectively. Banda Sea water outflows the In­

donesian Seas through the Ombai Strait and Timor Passage. This transport totals

14.9ST in August, 6.7 Sv in February, and 10.9ST annually, with the Timor Pas­

sage accounting for 63%, 27%, and 53% of this transport in August, February and

annually, respectively.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 35

250

200

150

100

50

50 100 150 200 250 i

Figure 2.4(d).

To disclose the influence of local winds consider experiments 8 (seasonally varying

transport without wind) and 2 (annual mean transport without wind). Total trans­

ports from experiment 8 for August and February and experiment 2 are shown in

Figures 2.4(d)-(f), respectively. Note that the specified open port transports already

contain the effect of seasonally varying global-scale wind, so turning off the local

winds can influence the distribution of the main flow between interior passages only.

First, we briefly describe properties of winds over the Indonesian Seas.

The southeast monsoon (SEM), winds out of the southeast, acts from July to

September while the northwest monsoon (NWM), winds out of the northwest, acts

from November to March. In August wind stresses on the order of 0.05 - 0.1 Am ” 2

are typical for the Banda, Flores, Seram, Maluku, and Halmahera Seas, and the

southern part of the Makassar Strait. Note that the northward component of wind

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 36

Figure 2.4(e).

stress during the SEM is rather strong in the Seram, Halmahera and Maluku Seas.

Approximately the same magnitudes of wind stress are observed in February when

the southward component is rather pronounced in the Seram, Halmahera and Maluku

Seas. The annual mean wind stresses in the Banda Sea and the southern part of the

Makassar Strait are in the same direction as the SEM though magnitudes are reduced

by a factor of two, while they are very weak in the Flores, Maluku, Halmahera, and

Sulawesi Seas, and the northern part of the Makassar Strait.

We consider the effect of the wind by comparing transports through the main

passages with and without wind in August and February. We note that transports

through the open ports do not depend on whether we have wind or not. Transports

through the exit passages, Lombok Strait, Ombai Strait and Timor Passage, are not

specified so we can expect that they will depend on the wind field. The results

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 37

50 100 150 200 250 j

Figure 2.4(f).

show that dependence on the wind field is weak. The same is true for transports

through other main passages. For example, turning the winds off in August results in

increased transport through the Makassar Strait of < 1ST and a decrease through the

Halmahera Sea of < 1 Sv, while in February transport through Makassar decreases by

~ 2Sv and increases by ~ 1 Sv each through the Maluku and Halmahera Seas. Figure

2.5 clearly presents the seasonal variations of transports through the main eight

passages. The annual harmonic is dominant in all transports except the Lifamatola

Passage and Ombai Strait. Note that amplitudes of transports vary for different

passages.

2.3.2. Surface Circulation

Surface circulation patterns for experiments 7 in August and February, and for

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5 6 7 8 9 10 11 12 5 6 7 8 9 10 11 12

> (0 5 4 2 4.5 5 6 7 8 9 10 11 12 5 6 7 8 9 10 11 12 time(years) time(years)

Figure 2.5. Transport time series through important passages and straits for four different model runs after 10 years of model run. The seasonally varying experiments with (exp 7) and without local winds (exp 8 ) are given by the solid and broken lines, respectively. The annual mean experiments with (exp 1) and without (exp 2) lo­ cal wind are given by the • and x symbols, respectively. The sections shown are: (a) North Makassar Strait, (b) Lombok Strait, (c) Lifamatola Passage, (d) Halma­ hera Sea, (e) Obi-Seram passage, (f) Mangole-Buru passage, (g) Timor passage, (h) Ombai Strait. Note the different transport scales on the y-axes. For exps 7 and 8 (lines), transports are calculated every 10 days while for exps 1 and 2 transports are calculated only at times given by • and x symbols.

experiment 1 are shown in Figures 2.6(a)-(c), respectively. In August an anticyclonic

gyre is seen in the Sulawesi Sea with the center displaced to the southeast. In the

northern Makassar Strait flow is predominantly southward with the velocity reaching

~ 40cm/s. In the south Makassar Strait most of the flow exits through the Lombok

Strait while the remainder enters the Flores Sea and then continues eastwards along

the Lesser Sunda Island Chain and finally exits the Indonesian Seas through the

Ombai Strait and Timor Passage. A westward current is seen throughout the Banda

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.6. Surface velocities in m/s. Vectors for experiment 7 for August and February, and experiment 1 are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 for August and February, and experiment 2 are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration.

Sea. A strong surface current from the NGCC port flows southward along the western

side of the Halmahera Sea and then travels directly west across the Seram Sea into

the Banda Sea, turning mostly southward at the southeastern tip of Sulawesi Island.

A branch of this current flows to the north basically through the Lifamatola Passage

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.6(b).

and Maluku Sea. Note also a northward current through the Seram-New Guinea and

Buru-Seram passages.

In February the center of the anticyclonic gyre in the Sulawesi Sea appears to

bedisplaced to the north, supplying a strong current into the northern Makassar

Strait. In the northern Makassar Strait the flow is everywhere to the south with clear

westward intensification; velocities in the east are ~ 50 cm/s and up to 100 cm/s

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.6(c).

in the west. In the south Makassar Strait the flow is everywhere to the east. This

eastward flow extends across the entire Banda Sea, to the Seram and Halmahera Seas

in the north and to the Timor Passage and the Lesser Sunda Island Chain in the

south. A rather narrow westward current is observed to the south of Burn Island.

Currents in the Maluku Sea and Lifamatola Passage as well as in the Seram-New

Guinea, and Buru-Seram passages are to the south.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 42

In the annual mean circulation (experiment 1) the center of the anticyclonic gyre

in the Sulawesi Sea is displaced similarly to February. In the northern Makassar

Strait flow is northward in the east and southward in the west with velocities of ~ 20

to 30cm/s in both directions. In the south Makassar Strait most of the flow enters

the Flores Sea, where an anticyclonic gyre is formed. Westward currents are seen

throughout the Banda and Seram Seas but a narrow eastward current is observed near

the Lesser Sunda Island Chain that exits the region through the Ombai Strait and

Timor Passage. Northward flow is seen in the Maluku Sea and Lifamatola Passage

as well as in the Seram-New Guinea, and Buru-Seram passages, similar to August

circulation. In the Halmahera Sea flow is southward in the west and northward in

the east.

Surface circulation patterns with no local wind for experiments 8 in August and

February and for experiment 2 are shown in Figures 2.6(d)-(f), respectively. The lo­

cal Ekman transport is 3 to 6m2s~l giving a wind-driven velocity of 0.05 to 0.1ms ' 1

in the upper 50m layer. Therefore the difference in surface velocities can be notice­

able. First we observe that velocity patterns with the local wind are more organized

compared to the corresponding patterns without the local wind. The influence of

local winds is clearly seen when comparing the directions of some surface currents in

experiments with and without wind.

In August we observe that surface currents are oppositely directed in experiments

7 and 8 (Figures 2.6(a) and (d)) in the southern part of the Sulawesi Sea, Maluku

Sea and Lifamatola Passage, northern part of the Banda Sea (to the south of the

Seram and Buru Islands), and the eastern part of the Makassar Strait (to the north

of the Labani Channel). These differences can be explained by the action of the local

wind stress. Invoking the geostrophic relations, we can determine the direction of

geostrophic surface currents in the above mentioned areas by estimating the direction

of SSH gradients (Figure 2.7(d)). It is seen that the y-gradient of SSH is positive

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.6(d).

in the Sulawesi Sea and to the south of Seram Island while the ./.-gradient of SSH

is positive in the Makassar Strait and Lifamatola Passage. Such gradients generate

exactly the surface currents shown in Figure 2.6(d). Taking into consideration the

direction of wind during the SEM we conclude that it is the SEM that generates the

differences in the direction of currents in Figures 2.6(a) and 2.6(d). Note that the

directions of gradients of SSH are essentially the same in Figures 2.7(a) (with wind)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.6(e).

and 2.7(d) (without wind).

In February surface currents are oppositely directed in experiments 7 and 8 (Fig­

ures 2.6(b) and (e)) in the northern part of the Makassar Strait (to the north of

the Labani Channel), in the Lifamatola Passage and Maluku Sea, and in the north­

ern part of the Banda Sea. The difference is also noticeable at the section between

Mangole and Buru Islands. The x-gradient of SSH is positive in the Makassar Strait

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 45

0.0 0.2 0.4 0.6 0.8 1.0 1.2

1 0 °N

5 ° N

0 °

5 ° S

1 0 ° S

1 5 ° S

2 0 ° S 115° E 120°E 125°E 130°E 135°E 140°E

Figure 2.6(f).

and negative in the Lifamatola Passage and at the Mangole-Buru section. The y-

gradient of SSH is negative in the northern part of the Banda Sea. It is easy to see

that the differences in the direction of these currents is generated by the NWM.

In the annual mean circulation (Figures 2.6(c) and (f)) the difference of direction

of currents is clearly seen throughout the Banda Sea. The annual mean wind is weak

and the situation resembles August more than February. Note that the velocity is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 46

reduced by a factor of two relative to August in the northern Makassar Strait while

the southward component is not as broad in the southern Makassar Strait.

2.3.3. Sea Surface Height

SSH for experiment 7 in August and February and experiment 1 are shown in

Figures 2.7(a)-(c), respectively. First, we compare absolute extremes. In August the

maximum SSH is located to the north of the Halmahera Sea (~ 16cm) while the

minimum is located to the south of the Lombok Strait (~ -20cm). In February the

maximum SSH is to the south of Sulawesi Island (~ 12cm) while the minimum is

to the south of the MC port (~ -24cm). The general structure of the annual mean

SSH is similar to the August pattern. Maximum SSH is located to the north of the

Halmahera Sea (~ 8 cm) and minimum SSH is located to the south of the MC port

(~ -17cm).

In August a high of ~ 8 cm is seen in the north of the Sulawesi Sea with a broad

low in the south. In the Makassar Strait a local maximum of ~ 0cm is seen in

the north with a local minimum of ~ —5cm in the south. In the Maluku Sea SSH

decreases from a maximum of ~ 2cm in the north to a minimum of ~ — 2cm at the

Lifamatola Passage. SSH in the Halmahera Sea also slopes downward from north

to south with a maximum of ~ 6cm in the north and a minimum of ~ 3cm in

the south. In the Banda Sea a local minimum of ~ - 8 cm is seen in the east and

the surface slopes upward across the full extent of the sea to the Flores Sea (local

maximum ~ 0cm). A relatively strong local gradient of SSH extends from the Labani

Channel along the southwestern stretch of Sulawesi Island, to the south Flores Sea

and running eastward along the Lesser Sunda Island Chain to the Ombai Strait and

Timor Passage.

In February, the SSH is essentially constant in the Sulawesi Sea to the west of

the Sangihe Ridge. The Makassar Strait slopes from a high of ~ 10cm in the south

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 50 1-0.15

50 100 150 200 250

Figure 2.7. Sea surface height. Values for experiment 7 for August and February, and experiment 1 are shown in (a) (above), (b) and (c), respectively. Values for experiment 8 for August and February, and experiment 2 are shown in (d), (e) and (f), respectively. These snapshots were taken after ten years of model integration.

down to a low of ~ 6cm in the north. The SSH both in the Maluku and Halmahera

Seas slopes weakly down from the south to the north from a high of ~ 5cm to a low

of ~ 3cm. A local high of ~ 10cm is seen in the easternmost Banda Sea, decreasing

to ~ 7 cm in the northern Banda and Flores Seas. This value decreases to a SSH of

~ 0cm in the south Flores and south Banda Seas.

The pattern in SSH for the mean annual experiment is similar to that seen in

August though not as pronounced; the low in the eastern Banda Sea is ~ - 2cm; the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 48

high in the Flores Sea is ~ 4cm; and the low in the south Makassar Strait is ~ Ocm.

The effect of local wind on SSH is generally not substantial (see Figures 2.7(d)-

(f)). In August the differences in SSH are still noticeable when the winds are removed.

SSH patterns in the Banda Sea are slightly changed with the value in the east in­

creasing to ~ -5cm. In February the pattern changes very little with a maximum

increase of ~ 2cm in the eastern Banda Sea. The changes in annual SSH are even

less noticeable.

2.4. Estimate of Terms in Heat and Salt Equations

Before estimating the role of the different terms in the heat and salt-diffusion

equations we clarify some important points. The equation for specific internal energy,

em, is formulated as (see, e.g., Kamenkovich 1977; p. 41)

9 / ^ ( 9X 1 3 W7 (pem) + div pemu + q + = X Paf3ea/3 , (2-7) ot 8 S ) rp / T,p a ,0 = 1 where p is the density, u is the velocity of a fluid particle, q is the density-of-heat-flux

vector, I is the density-of-salt-mass-flux vector, Xm is the specific enthalpy, T is the

temperature in Celsius scale, s is the salinity (as a non-dimensional ratio), paj3 is the

stress tensor, and eap is the rate-of-strain tensor.

The last term in the brackets describes the influence of salt diffusion on the

change of the specific internal energy while the term 011 the right-hand side of the

equation describes the mutual conversion of mechanical and internal energy. Both of

these terms are usually neglected. So, the rate of change of internal energy per unit

volume, pem, is controlled by the advection of this property and the divergence of the

density-of-heat-flux vector. The specific internal energy is defined with respect to

an undetermined reference state, thereby influencing the advection of this property

through an open surface.

Following Warren (1999), we will consider the anomaly of the specific internal

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 49

50 100 150 200 250 i

Figure 2.7(b).

energy, e'm(T,s,p) = em(T,s,p) — e(0,0.035, p0), where p is the pressure, and p0

is the mean atmospheric pressure. Due to the mass conservation equation we can

replace em in (2.7) by e'm. Then the advection of pe'm through any surface can be

unambigously determined. It is important that Warren (1999) has shown that with

sufficient accuracy

e'm = Cp 0 , (2.8)

where the specific heat capacity at constant pressure, cp, can be considered as con­

stant (cp = 4 x 103Jkg~1), and 0 is the potential temperature in Celsius scale.

Replacing the density, p, by the mean density, p0, we rewite (2.7) as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 50

Figure 2.7(c).

~ + div (du + = 0. (2.9) dt \ pocpJ

Equation (2.9) is usually called the heat equation. Substituting the turbulent density-

of-heat-flux vector for the molecular density-of-heat-flux vector q and using the POM

parameterization gives the POM heat equation. Integrating this equation, first over

a from -1 to 0, and then over a fixed horizontal area E, and using the boundary

conditions for wpom (at the surface and bottom) and for 9 (at the bottom) gives

^-tjD9dE + fDT(^i)dT - I ( ~ ~ ) dS + | DrW(0-edim)dZ E r E ^ E

— H A h — (6 — 9dim) dT = 0, (2.10) r

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 51

Figure 2.7(d).

where the bar implies the integral over a from -1 to 0; T is the boundary contour

of E; n is the outward normal vector to this contour; subscript 0 indicates that the

expression is calculated at the sea surface (a = 0); is the nudging coefficient

dependent on a; 9ciim is the climatology of 9; AH is the coefficient of horizontal d turbulent diffusion; and 7 — is the derivative in the n direction; D is the total depth, on D = H + rj, where 77 is the sea-surface height; and KH is the coefficient of vertical

turbulent diffusion. The expression for the last term on the left-hand side of (2.10)

follows from the parameterization used in the POM (see Mellor 2004; formulas (11)

and ( 12)).

In equation (2.10) we have a balance between (from left to right) the rate of time

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 52

50 100 150 200 250 i

Figure 2.7(e).

variation of the volume integrated 0 (or S)\ the sum of horizontal advective fluxes

of 6 (or S) through the side boundaries of the volume; the vertical flux of heat (or

salt mass) through the sea surface integrated over the sea-surface side of the volume

(with the sign changed); the nudging term integrated over the volume; and finally

the sum of horizontal diffusive fluxes through the side boundaries of the volume (with

the sign changed).

We now break the area down into boxes and consider equation (2.10) and a similar

equation for salinity S for these boxes. All terms in (2.10) are calculated directly

from the simulated fields of the horizontal components of velocity (ms-1), potential

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 2.7(f).

temperature (°C) or salinity, and coefficients of vertical and horizontal diffusion (all

data are for August). The results for the most important boxes are presented in

Tables 3 and C.l - C.3 in Appendix C. The boxes considered are shown in Figure

2. 8.

The sum of horizontal advective fluxes through all sides of a given box, vertical

flux at the sea surface, nudging term, and rate of time variation appear to be basically

of the same order. However, the sum of horizontal diffusive fluxes through all sides

of the box is at least two orders of magnitude smaller than all other terms. Note, the

sum of horizontal advective fluxes is several orders of magnitude smaller than the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 54

calculated values 9 S rate of time variation (xlO8) -8.9xl0 -3 1.3xl0 - 3 horizontal advective fluxes (xlO8) Lifarnatola passage (j=152), Q=-2.7Sv -7.6xl(T 2 -9.3xl0 _1 Between Mangole and Buru (i== 130), Q=6.3Sv 1.2 2.2 Between Buru and Seram (j=134), Q=0.6Sv 1.2 xlO_I 1.9xl0 -1 Between Obi and Seram (i=145), Q=-4.2Sv -1.2 -1.5 Sum of horizontal advective fluxes 6.8 xlO - 3 -5.7xl0 ~4 horizontal diffusive fluxes (xlO8) Lifarnatola straits (j = 152) -1.6 x 10-4 1.4xl0 - 6 Between Mangole and Buru (i=130) 1.9xl0 -4 -1.2 x l 0 -6 Between Buru and Seram (j=134) ~3.3xl0 - 5 2 .8 x 10-7 Between Obi and Seram (i=145) -2.9xl0 “ 5 -5.0xl0 “ 7 Sum of horizontal diffusive fluxes -3.4xl0 - 5 l.lx lO -6 vertical flux at the surface (xlO8) -2.3xl0 - 3 -8.4x 10 -3 nudging (xlO8) -2.9x 10-3 7.7xl0 “ 3 sum of all terms (xlO8) -7.2 xlO -3 -4.8 xlO - 5

Table 3. Estimates of different terms in the depth integrated heat, (9), and salt diffusion, (S'), equations for the Seram Sea box. August values are given for the seasonally varying experiment with local wind. Qs are the total transports through the sides of the box. Negative values of Q are into the box and positive values are out of the box. The last line gives the estimate of accuracy of calculations. Values for specific heat and reference density are, respectively, cp = 4 x 103J o C~1kg~1, Po = 1025kgm~3.

leading advective flux through a side of the box, while the sum of horizontal diffusive

fluxes is always on the order of the diffusive flux through one side of the box. It

is well known that the POM parameterization of horizontal diffusion can generate

some false dynamical effects. These effects, if they are present at all, are very small

in our model.

Tables 3 and C.l - C.3 (Appendix C) show these balances in more detail. For the

Seram box (Table 3) the sum of advective fluxes, vertical diffusion at the sea

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 55

50 100 150 200 250 i

Figure 2.8. The eight boxes for which estimates of terms in the heat and salt diffusion equations have been made. The boxes are: 1. Seram Sea, 2. Banda Sea, 3. Makassar Strait, 4. Lombok, 5. East Maluku Sea, 6 . Karakelong-Sangihe-Morotai, 7. Halmahera Sea, and 8 . West Sulawesi Sea.

surface, nudging, rate of time variation, and the sum of all terms (residual) are on

the same order (for 9); for S, the balance is essentially between vertical diffusion

and nudging while the residual is two orders of magnitude smaller. For the Banda

box (Table C. 1) the nudging and rate of change of time variation balance each

other while the residual is one order of magnitude smaller for 9; for S the balance is

essentially between the sum of advective fluxes, vertical diffusion and nudging while

the residual is two orders of magnitude smaller. For the Makassar box (Table C.

2) the balance is essentially between the sum of advective fluxes, nudging, and the

rate of time variation while the residual is on the order of the nudging term (for

9); while for S, the dominant terms are the vertical diffusion and nudging while the

residual is on the order of the vertical diffusion term. For the Lombok box (Table C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 56

3) all four terms - the sum of advective fluxes, vertical diffusion at the sea surface,

nudging, rate of time variation, and the residual are on the same order ( 6); for S,

the vertical diffusion, nudging, and the rate of time variation are on the same order

while the residual is one order of magnitude smaller. Similar relations hold for other

boxes (East Maluku, Karakelong-Sangihe-Morotai, Halmahera, and West Sulawesi).

Note, that the sum of all terms in (2.10) (residual) is sometimes on the order of these

terms rather than at least one order of magnitude smaller. This is seemingly due to

the method of calculation of horizontal advective fluxes in (2.10). Running the POM

we used the Smolarkiewicz iterative scheme for the advection terms in the 6- and

5-equations, while in estimating the horizontal advection in the integrated forms of

these equations we applied the corresponding expressions in ( 2.10) directly.

2.5. Comparison with Observations

The comparison of simulated and observed basic structures of the surface circu­

lation and sea-surface height patterns along with magnitudes of transports through

main passages appears to be satisfactory. We now consider in detail the observed

data.

2.5.1. Transports and Surface Circulation

Molcard et al. (1996) estimated the transport through the Timor Passage with

data from two current moorings deployed near the southwesternmost tip of Timor

Island between March 1992 and April 1993. They also made continuous CTD casts

across the passage for 2 days during deployment. Their estimate of the mean trans­

port for this period is between 3.4 and 5.3 ± l.bSv. In the upper 500m, the transport

is at a minimum of 1 — 2 Sv during the NWM and at a maximum of 3 — ASv during the

SEM. They found a mean surface velocity of ~ 40cm/.s directed toward the Indian

Ocean along the axis of the passage. Surface velocity is weakest but in the same

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 57

direction during the NWM. Model transport through the Timor Passage is at a min­

imum of 2 Sv during the NWM and increases to almost lOSu during the SEM (Figure

2.5(g)). Model surface velocities are maximal during the SEM (Figure 2.6(a)), mini­

mal during the NWM (Figure 2.6(b)) and the mean values are in agreement with the

calculation of Molcard et al. (1996) (Figure 2.6(c)). Molcard et al. (1994) also cal­

culated the transport through the Timor Passage between August 1989 and August

1990 from current meter moorings and Indonesian coastal tidal gauges. They calcu­

lated a mean transport for the period of 7 Sv. In early September 1989 the transport

was as high as IlSu, supplying at least half of the total transport of 22Sv between

Bali and Australia (see Fieux et al. 1994). Molcard et al. (1994) conclude that

the Timor Passage may supply more than 50% of the Indonesian Throughflow (ITF)

during the SEM. The model gives a maximum transport through the Timor Passage

of almost lOSu during the SEM which is more than half of the total transport of the

ITF (17ST) at this time (Figure 2.4(a)).

Cresswell et al. (1993) also estimated a mean transport through the Timor Pas­

sage of 7Sv from shipboard ADCP measurements made in October 1987 and March

1988 and from two current moorings deployed between March and December 1988.

An analysis of drifters show that surface velocities exceeded 1 m /s in July 1983.

These observations are in good agreement with model results (see Figure 2.5(g);

Figure 2.6(a)).

Molcard et al. (2001) calculated the mean transport through the Ombai Strait

from a current mooring deployed there for one year in December 1995. They found

a transport of 5 ± 1ST with a maximum of 6Sv in August and a minimum of 4 Sv.

Maximum surface velocities of ~ 60cm /s are seen in August. The model gives a mean

transport of ~ 5.5 Sv (Figure 2.4(c); Figure 2.5(h)), with a maximum in August of

~ 5.1 Sv (Figure 2.4(a); Figure 2.5(h)). Model surface velocities reach a westward

maximum during August as well (Figure 2.6(a)).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 58

In the Lombok Strait, Murray & Arief (1988) observed a distinct annual cycle in

transport from current meter moorings, deployed between January 1985 and January

1986. They observed a minimum transport of ~ 1 Sv from February to May, a

maximum of ~ 4.05T in the northern summer months and a mean value of 1.7 Sv

for the duration of the deployment. The model gives a similar cycle (Figure 2.5(b))

though maximum transports do not exceed 2.2 Sv. From the same data set, Murray

et al. (1990) observed southward surface velocities throughout the year with brief

periods of reversal towards the north. Model surface currents are always to the south.

Hautala et al. (2001) estimated mean transports through the exit passages of the

ITF from shallow pressure gauge arrays (SPGA) deployed in the passages between

December 1995 and May 1999 and from ADCP surveys during original and subse­

quent deployments and recovery. They estimated mean total transports for the two

year period (1996 - 1997) of 2.6 ± O.SSv through the Lombok Strait; 2.6 ± 0.8ST

through the Ombai Strait; and 3.2 ± 1.8ST through the Timor Passage. These esti­

mates compare reasonably well with model results (Figure 2.5 (b, h, g)).

Susanto & Gordon (2005) estimated total transport through the Makassar Strait

with data from two current meter moorings including near surface ADCP measure­

ments deployed in the Labani Channel between December 1996 and July 1998. They

found a mean transport of 8.1 ± l.5Sv for the entire period and 7.9 ± 1.2Sv for

the 1997 year. The results represent an improvement on the previous estimates

of Gordon & Susanto (1999), Gordon et al. (1999) and Wajsowicz et al. (2003),

since those estimates were hindered by the absence of a surface layer record from

the ADCP. For the same data set, Gordon & Susanto (1999) and Gordon et al.

(1999) estimated transports of 9.2 and 9.3 Sv, respectively, from four simple velocity

profiles (maximum, minimum, average, and seasonal dependence). Wajsowicz et al.

(2003) estimated a mean transport of 6.4 Sv using a normal mode approach based on

monthly data; while Vranes et al. (2002) estimated a mean transport of 8 ,9Sv using

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 59

data only from the upper 700m of the westernmost mooring. Maximum estimates of

transport briefly exceeded 14 Sv in June - July 1997 and August 1998 while minimum

transports dipped below 4Sv in December - January 1997 and during the NWM of

1998. All of these estimates are in good agreement with the model results (see Figure

2.5(a)).

Wajsowicz et al. (2003) also found maximum and minimum transport values of

16.0 and 4.7Sv, respectively. The ADCPs lasted only 3 months on the westernmost

mooring and 8 months on the easternmost. Nevertheless, surface velocity values were

extrapolated from the ADCP data and the shallowest current meter for the entire

deployment period. Comparing ADCP observations between December 1996 and

February 1997 yields mean surface velocities of 35 cm/s southward for the western­

most mooring, and of 22 cm/s for the easternmost. This appears to be the result

of the western boundary effect and is also produced b}^ the model (see, e.g., Figure

2.6(b)). Mean surface southward velocities of ~ 25 cm/s at the Labani Channel are

in good agreement with model results (Figure 2.6(c)).

Van Aken et al. (1988) estimated a deep transport of ~ 1.5SV across the Lifam-

atola Passage from a current mooring deployed between January and March 1985,

in good agreement with model results (see, e.g., Fig 2.5(c)).

Wyrtki (1961) was the first to give a comprehensive description of the oceanogra­

phy of the Indonesian Seas and of the entire Southeast Asian region. Using both his

and historic observations, including data from merchant ships (see Wyrtki (1960) for

further discussion), he found a high correlation between the monsoon winds and the

ocean circulation. Just as monsoon winds practically reverse direction twice a year,

the ocean currents are also reversed in large areas of the region. W yrtki’s (1961) de­

scription of the surface circulation is as follows. In February the MC extends to the

southern tip of Mindanao Island where it splits into two parts. The larger portion

continues southward to the Morotai Basin before turning sharply to the east, form­

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 60

ing the cyclonic Mindanao eddy and the core of the NECC. The remainder enters

the Sulawesi Sea, though this is partially returned toward the NECC as a portion of

the broader Mindanao eddy retroflection. This circulation pattern essentially per­

sists throughout the year though the extent and location of the Mindanao eddy may

change. The area to the north of New Guinea is marked by the near surface seasonal

variations of the NGCC. This area is influenced by the southeast trade winds from

May until October and by a branch of the northeast trade winds between November

and April. As a result, the surface currents are strongest out of the southeast in

August, when the Halmahera eddy is developed, and strongest toward the southeast

in February, when the Halmahera eddy is not developed.

Wyrtki (1961) showed also that in February part of the MC crosses the central

Sulawesi Sea to enter the Makassar Strait. In addition, there is an anticyclonic cell

to the north, and cyclonic cell to the south. This system of circulation is supported

from February to September. The Mindanao eddy shifts to the east in October and

no surface flow from the Sulawesi Sea enters the Makassar Strait.

In the Makassar Strait the surface flow is directed to the south throughout the

year with small velocities as a rule. Currents are strongest in February and March,

and weakest from July to September. In the south of the Makassar Strait, waters

leaving the strait flow to the west during the SEM and to the east during the NWM.

In the Flores Sea, Wyrtki (1961) described the surface flow as being uniform

and eastward from December to March, when the winds are westerly, and westward

for the remainder of the year, when the winds are southeasterly. He noted that a

weak eastward current just to the north of the Lesser Sunda Island Chain persists

throughout the year.

Mean surface flow across the Lifarnatola Passage is southward at this time. This

current system is reversed during the SEM when surface waters from the Halmahera

Sea enter the Seram Sea and the Banda Sea before reaching the Flores Sea and mean

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 61

surface flow across the Lifarnatola Passage is to the north.

A southwestward current prevails throughout the year in the Timor Passage with

its axis running close to Timor. The current extends to Australia during the SEM

though velocities decrease across the passage. The current retreats toward Timor

during the NWM and a northeastward current forms in the south. During the NWM

the Timor current is supplied by the current flowing to the east along the north coast

of the Lesser Sunda Island Chain. During the SEAI the Timor current is supplied by

upwelling in the eastern Banda Sea.

The model fairly well reproduces the circulation patterns described by Wyrtki

(1961).

2.5.2. Sea Surface Height

Based on sea-level observations, Wyrtki (1961) concluded that the SSH of the

Banda Sea is intimately linked to the southeast Asian monsoon system. During the

NWM, surface water enters the Banda Sea from the west resulting in accumulation

of surface water there and raising of SSH. During the SEAI, this surface water enters

the Indian Ocean as part of the ITF. Wyrtki (1961) estimated the range in SSH in

the Banda Sea at ~ 10 — 20cm. Gordon & Susanto (2001) investigated surface-layer

divergence and sea level anomaly (SLA) in the Banda Sea. They calculated a time

series of SLA from the combined TOPEX POSEIDON/ERS altimeter data for the

period December 1992 to December 1999. An annual range of ~ 20cm was found

with a maximum of ~ 10cm in February and a minimum of ~ — 10cm in August.

The model is in good agreement with both Wyrtki (1961) and Gordon & Susanto

(2001), having a similar range of SSH between these months with a maximum in

February of ~ 10cm and a minimum in August of ~ 8 cm (see Figures 2.7(b) and

2.7(a)).

Bray et al. (1996) investigated large-scale sea level variations in the Indonesian

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 62

Throughflow region. They used data from the Tropical Ocean-Global Atmosphere

(TOGA) sea level archive of stations corrected by barometric pressure data from the

Comprehensive Ocean-Atmosphere Data Set (COADS). Bray et al. (1996) found

that most of the variability in SSH is contained in the annual signal. Annual values

of variability were least in the northern Indonesian Seas with a minimum amplitude

of ~ 2cm in the Sulawesi and Maluku Seas as well as in the Makassar Strait. The

amplitude increases to the south and east to a value of ~ 4cm in the Flores and

northern Banda Seas, to ~ 8 cm in the Timor passage and to a maximum value of

20cm in the southeastern Banda Sea. Maximum SSH in the Banda Sea region occurs

in January-February. SSH maxima occur later in the year in the northern Indonesian

Seas with a maximum occurring in the Sulawesi Sea several months later. Model SSH

is in good agreement with these observations. Model SSH annual amplitude is least

in the north with a minimum value of ~ 2cm in the Sulawesi Sea region, increasing

to a maximum value of ~ 20cm in the southeast Banda Sea.

Waworuntu et al. (2001) investigated the dynamics of the Makassar Strait by

analyzing inverted echo sounders (IES) and bottom pressure data (PIES) for the pe­

riod from October 1996 to March 1998 combined with TOPEX/POSEIDON satellite

derived sea height anomaly for the period October 1992 to April 1998. Both the

PIES and TOPEX/POSEIDON data show an annual signal in SSH in the Makassar

Strait with maximum SSH during the NWM and minimum SSH during the SEM.

TOPEX/POSEIDON data show greater annual amplitude of ~ 20cm in the north­

ern Makassar Strait, reducing to ~ 10cm in the southern Makassar Strait. Model

results are in good agreement with maximum values of SSH in the Makassar Strait

during the NWM and minimum values of SSH during the SEM. Annual simulated

amplitudes are ~ 10cm.

2.6. Summary and Conclusions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 63

The basic features of a developed high-resolution regional model of the Indone­

sian Seas circulation have been discussed. The model, based on the Princeton Ocean

Model (POM), has horizontal resolution of ~ lOfcra and 29 a-levels in the vertical. It

is well known that the regional dynamics of the Indonesian Seas is strongly controlled

by peculiarities of the bottom topography. Thus the primary motivation for devel­

oping the model is to take into account all basic features of the bottom topography

in the region. It is demonstrated that by using a horizontal resolution of 10km we

have successfully incorporated all important topographic features into this model.

Although the limitations of regional models are known, such models are useful in

gaining insight into the regional dynamics. We feel confident that important ques­

tions such as the partition of the main flow between passages and interaction of NPW

and SPW, the influence of local winds, overall momentum and energy balances, and

properties of turbulent mixing in the region can be successfully addressed with the

developed model.

The model has 250 x 250 cells in the horizontal and four open ports to sim­

ulate major currents into and out of the domain: the MC (Mindanao Current)

port through which North Pacific Water (NPW) enters the region; the NGCC (New

Guinea Coastal Current) port through which South Pacific Water (SPW) enters the

region; the NECC (North Equatorial Counter Current) port through which part of

both NPW and SPW leave the region; and the 10 (Indian Ocean) port through

which the Indonesian Throughflow (ITF) enters the Indian Ocean. Model topog­

raphy, interpolated from the ET0P05 database, has been edited and smoothed to

satisfy the POM recommendation for maximal permissible slopes, thereby decreasing

possible effects of the pressure gradient representation in the POM. The smoothed

topography has been thoroughly checked to ensure that all essential details have been

retained.

At the open ports, the normal velocity for the 2D (depth averaged) motion is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 64

specified based on the prescribed total transport through the port. The tangential

2D velocity is set to zero. For the 3D motion, Orlanski’s condition for the normal

3D velocity with nudging to the observed normal port velocity has been employed.

Different nudging coefficients for the inflow and outflow ports were used. The tan­

gential 3D velocity is set to zero. The boundary conditions for potential temperature

and salinity are similar to the 3D normal velocity condition. The turbulence energy

and length scale are set to zero. The remaining boundary conditions are standard

for the POM.

The POM basic equations have been slightly modified both in the port channels

and in the interior domain. The open port channels (5 cells in length) were intro­

duced to allow for tapering off of nudging effects outside the interior of the model

domain. For the momentum equations, friction and nudging terms have been added

in the inflow channels (MC and NGCC) and frictional terms only have been added

in the outflow channels (10 and NECC). For the temperature and salinity equa­

tions, nudging terms have been added at all ports. Nudging and friction coefficients

decrease toward the interior domain for all ports. In the main domain, additional

friction has been added to the momentum equations at some sills and passages. This

measure was used to take into account some specific processes, e.g., tidal effects, not

included in the model. Temperature and salinity have been nudged to climatology

near the surface to control the Mellor-Yamada parameterization of vertical turbulent

fluxes in the layer of strong salinity extrema. Weak nudging was included below

the thermocline to reduce the effect of the POM horizontal diffusion. Advection of

temperature and salinity has been modeled with two iterations of the Smolarkiewicz

scheme.

At the open ports normal velocities and total transports are specified from pub­

lished observations and some modeling results. At the 10 port we have basically

used the observed data of Fieux et al. (1994) along with other observations at the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 65

exit passages of the Indonesian Seas region (Lombok Strait; Ombai Strait; Timor

Passage). We have chosen transport values of 17Sv in August, 8 Sv in February and

an annual mean value of 12.5ST. At the NGCC port we have basically used the

observations of Ueki et al. (2003). We have chosen transport values of 28 Sv in Au­

gust, 17 Sv in February and 22.5 Sv for the annual mean. At the NECC port we have

basically used the observations of Qiu & Joyce (1992). We have chosen transports

of 46ST in August, 35.ST in February and an annual mean value of 40SV. At the

MC port we have basically used the observations of Lukas et al. (1991). We have

chosen transports of 35 Sv in August, 25 Sv in February and a mean annual transport

of 30Sv. Note that the port normal velocity profiles at the 10, NECC and MC ports

are assumed constant throughout the year, while the port normal velocity profiles

at the NGCC port are supposed different for August and February. All port normal

velocity profiles are scaled by the specified total transports. All seasonally varying

external factors (port transports, port normal velocities, local wind stress, tempera­

ture and salinity climatology) were approximated in time by the annual mean value

plus two Fourier harmonics.

The four experiments discussed in this paper - annual mean with and without

wind and seasonally varying with and without wind - were run for at least 12 years.

All of the experiments were fully spun up after 10 years.

The basic properties of simulated total transports through the main passages in

the region, surface circulation and sea-surface heights (SSH) have been discussed.

According to calculated transports, NPW splits into three branches upon entering

the model domain through the MC port. The first branch takes the western route

over the Sangihe Ridge, across the Sulawesi Sea and along the Makassar Strait before

diverging at the southern part of the Makassar Strait. The minor part of the flow

exits through the Lombok Strait while the major part enters the Banda Sea through

the Flores Sea. The second branch of NPW takes the north-south route through

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 66

the Morotai basin to the Maluku Sea, and crosses the Lifarnatola Passage into the

Seram Sea before finally entering the Banda Sea. The third branch of NPW exits

through the NECC port. SPW splits into two branches upon entering the model

domain through the NGCC port. The first branch takes the eastern route through

the Halmahera Sea into the Seram Sea where it encounters NPW on the north-south

route. The second branch exits through the NECC port. Banda Sea water exits the

Indonesian Seas through the Ombai Strait and Timor Passage. Transports through

the Makassar Strait, Maluku Sea, Halmahera Sea, Lombok and Ombai Straits, and

the Timor Passage are stronger in August and weaker in February relative to the

annual mean. The portion of NPW entering the Indonesian Seas relative to that

leaving through the NECC port is fairly constant throughout the year. Most of

this water takes the western route. The portion of SPW entering the Halmahera

Sea compared to that exiting through the NECC port varies considerably with the

seasons. Almost all of the NPW taking the north-south route through the Maluku Sea

and SPW crossing the Halmahera Sea enters the Banda Sea through the Mangole-

Buru Passage. The total transport through Mangole-Buru is about 2/3 of that

through the Makassar Strait in August and the mean annual experiment, and about

1/2 that through Makassar in February. The flow through Lombok is fairly constant

at about 1./4 of the total Makassar transport with the remainder entering the Banda

Sea. At the exit passages, the relative changes in transport with the seasons are

strongest for the Timor Passage, moderate for the Lombok Strait and insignificant

for the Ombai Strait. Turning off the local winds does not substantially influence the

transport through the main passages in the model domain. For almost all passages

the annual harmonic appears dominant with substantially different amplitudes.

Surface circulation patterns change substantially with the seasons. In August the

surface current is everywhere to the west in the interior of the Indonesian Seas in

response to the south-east monsoon (SEM) with the exception of a narrow eastward

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 67

band along the Lesser Sunda Island Chain. In February the flow is everywhere to

the east in the Banda Sea. Flow in the Makassar Strait is to the south with clear

westward intensification seen to the north of Labani Channel. For the mean annual

experiment, flow is to the west in the Banda Sea due to the rather strong mean

local wind there. Turning off the local wind reveals a noticeable change in surface

circulation. First, the surface circulation patterns are not as organized. Surface

circulation patterns are oppositely directed for large areas of the region. Invoking

the geostrophic relations shows that differences in current velocities are generated

by the local wind stress.

SSH values are very reasonable throughout the model domain. Values are affected

by seasonally changing transports and local wind stresses. In the Banda Sea region,

the SSH slopes up from the east to the west during the SEM and down from the east

to the west during the north-west monsoon (NWM). In the annual mean the SSH is

similar to August though amplitudes are reduced due to the decrease in local wind

stress and transports. SSH amplitudes change when the local winds are turned off

but the effect is not as substantial as the seasonal one.

The role of different terms in the heat and salt equations was investigated by

dividing the region into a number of boxes. For any given box, the sum of the

horizontal advective fluxes of temperature (salinity) through all sides of the box is

on the same order as the vertical heat (salt) flux at the surface, nudging term, and

the rate of time variation of the box integrated temperature (salinity). The sum of

horizontal diffusive fluxes of temperature (salinity) through all sides of a given box

are two orders of magnitude smaller than all other terms. The leading advective

flux term for a given box is several orders of magnitude greater than the sum of

the advective fluxes through the box while the sum of the horizontal diffusive fluxes

through all the sides of the box is always on the same order as the flux through one

side. It is well known that the POM parameterization of horizontal diffusion can

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 68

generate false dynamical effects. These effects, if they are present at all, are very

small in our model.

The comparison of the basic structure of the model surface circulation, SSH

and total transport values through the main passages with observations appears

satisfactory. At the exit passages (Timor Passage, Ombai Strait and Lombok Strait),

seasonal variabilities of transport and surface velocities in the model are in good

agreement with available data. Model values of annual mean transport through these

exit passages are also in good agreement with transports estimated from observations.

Model transports in the Makassar Strait are in good agreement with the results of

several investigators. The western boundary effect seen in the model Makassar Strait

was also observed in ADCP data. The model transport at the Lifarnatola Passage

is in reasonable agreement with available observations. Details of the model surface

circulation are in reasonable agreement with Wyrtki’s (1961) description. Model

results in the region show that the SSH is modulated by the monsoons. This result

is verified by observations. Observations in the Banda Sea show SSH values in the

east are least during the SEM and greatest during the NWM. SSH slopes up toward

the west in August and down to the west in February. This pattern is reproduced

by the model and the amplitudes are of the same scale. Similar agreement is found

in the Makassar Strait.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 69

CHAPTER 3

THE INFLUENCE OF BOTTOM TOPOGRAPHY ON TEMPERATURE

AND SALINITY DISTRIBUTIONS

3.1. Introduction

One of the most important features of the Indonesian Seas circulation is the

transport of mass, internal energy (heat), and salt through the region from the Pacific

to the Indian Ocean. The estimate of these transports depends on the distribution

of potential temperature, salinity and velocity at the entering and exiting passages

of the region. The numerical model and properties of simulated velocity fields have

been discussed in Chapter 2. The main objective of this chapter is to describe the

basic properties of the simulated potential temperature and salinity distributions

and identify physical processes responsible for their formation. The simulated and

observed properties of these distributions will be compared to validate the model.

An outstanding feature of the Indonesian Seas is the complexity of the bottom

topography. This topography has a profound effect on the distribution of tempera­

ture and salinity in the region. In setting up the model we have taken into account

all important topographic features by using high horizontal resolution and relevant

smoothing of the bottom relief (see Chapter 2). There are two entering types of

water recognized basically by their vertical salinity structure. North Pacific Water

(NPW) entering the Indonesian Seas region as a branch of the Mindanao Current is

characterized by a surface value of ~ 34.10; a salinity maximum of ~ 34.74 at a depth

of ~ 150m; a deep minimum of ~ 34.46 at a depth of ~ 350m; and a deep value

of ~ 34.66. South Pacific Water (SPW) entering the Indonesian Seas as a branch of the New Guinea Coastal Current is characterized by a surface value of ~ 34.15;

a strong maximum of ~ 35.50 at a depth of ~ 100 — 150m; a weak minimum of

~ 34.61 at a depth of ~ 1500m; and a deep value of ~ 34.65. The temperature

and salinity characteristics of NPW and SPW change as waters flow through the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 70

Indonesian Seas region. What factors control these changes? In general, the water

properties are advected and diffused. We will show that below 500m the advection is

totally controlled by topographic features in the region. Large elevations of bottom

relief break the region down to separate basins having different temperature and

salinity stratifications. These topographic elevations impede the advection of cold

and salty water from a basin (located upstream) to the neighboring basin (located

downstream). We will consider temperature and salinity sections along the paths of

NPW and SPW. By comparing upstream and downstream temperature and salinity

distributions for each substantial topographic elevation we will identify the impact

of the topographic feature. Finally, we will choose salinity as an important tracer

and estimate the salinity advection through the sections located across the main flow

to see what salinities are really advected into the basin.

There are two paths of flow of NPW and one path of flow of SPW to the Banda

Sea from where waters leave the region, essentially through the Ombai Strait and the

Timor Passage. The first route of NPW (the western route) is through the Sulawesi

Sea. Makassar Strait arid Flores Sea. Along this route, the rather shallow Dewakang

Sill of depth ~ 600m is encountered, blocking the flow of deep saline water towards

the Flores Sea and thereby the Banda Sea. We will show that in the upper ocean

(500m) the Indonesian Throughflow primarily consists of NPW taking the western

route through the Makassar Strait and into the Banda Sea. The second route of

NPW (the northern route) is through the Morotai basin, Lifarnatola Passage and

Seram Sea. We will show that basically the deep salinity of the Banda Sea is formed

through this route. We will also discuss the long standing question regarding the

role of SPW (see, e.g., Aken et al. 1988; Godfrey et al. 1993; Gordon 1995; Godfrey

and Wilkin 1995; Gordon & Fine 1996; Gordon et al. 2003a; and references therein).

Essentially, SPW enters the region through the Halmahera Sea. We will show that

the northern and southern Halmahera Sills completely block the flow of deep SPW

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 71

into the Indonesian Seas. But in the upper layer (0 — 500m) some SPW can probably

penetrate to the Maluku and Seram Seas to mix there with NPW and finally reach

the Banda Sea.

In Section 3.2 we will estimate the simulated transports of mass, internal energy

(heat), and salt through the Indonesian Seas region and contributions of deep parts

of the seas to these transports. In Section 3.3 we will analyze the August simulated

temperature and salinity distributions in the region and processes that form these

distributions. In Section 3.4 we will describe the seasonal variations of the simulated

distributions. In Section 3.5 a detailed comparison of simulated and observed char­

acteristics will be provided to validate the model. Finally, in Section 3.6 the basic

results of the study are summarized.

3.2. Transports of Mass, Internal Energy and Salt

We begin with the calculation of transports of mass, internal energy (heat), and

salt through the Ombai Strait, Timor Passage, and Lombok Strait along with trans­

ports through the Mindanao Current (MC), New Guinea Coastal Current (NGCC),

and North Equatorial Counter Current (NECC) ports. The results for August and

February are presented in Table 4. In general, the values of total transports are

in agreement with published simulated and observed data. A short review of this

comparison is given in Section 3.5.

Analyzing Table 4, we notice first the substantial seasonal changes of transports

caused by seasonally varying water transports through the MC, NGCC, and NECC

ports and by local winds. Second, transports of internal energy (heat) and salt in the deep parts (below 500m) of the Ombai Strait and Timor Passage appear rather

substantial. Therefore, the analysis of processes responsible for the formation of

distribution of potential temperature and salinity in the deep parts of Indonesian

Seas (especially in the Banda Sea) is very important. Such an analysis will be

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 72

mass transp heat transp salt transp ( x l 09kg/s) (PW) ( x l 08 kg/s) Aug Feb Aug Feb Aug Feb Ombai (i=90) total 5.55 5.06 0.31 0.18 1.91 1.75 0-500m 3.07 1.79 0.26 0.12 1.05 0.62 Timor (i=90) total 9.55 1.94 0.69 0.30 3.28 0.67 0-500m 9.75 4.23 0.69 0.35 3.35 1.46 Lombok (j=110) total 2.18 1.36 0.19 0.12 0.75 0.47 0-500m 3.14 1.77 0.21 0.13 1.08 0.61 MC port (j=246) total -35.70 -25.80 - 2.22 -1.51 -12.32 -8.91. 0-500m -30.90 -21.71 -2.19 -1.40 -10.66 -7.50 NGCC port (i=246) total -28.51 -17.62 -2.02 -0.98 -9.97 -6.17 0-500m -26.93 -16.16 -1.96 -0.93 -9.42 -5.67 NECC port (i=246) total 46.94 35.06 3.85 2.83 16.25 12.19 0-500m 44.97 34.28 3.76 2.79 15.57 11.92 Table 4. The simulated transports of mass, internal energy (heat), and salt through MC, NECC, and NGCC ports, along with Lombok and Ombai Straits, and Timor Passage for August and February. Additionally, the corresponding transports for the upper 500m are given. The specific heat capacity at constant pressure is cp = 4 x 103 J °C~l kg“T The mean density is given by p0 = I025kgm~3, IPW = 1015 Js~l.

presented in the next section. We remind readers that the Banda Sea is the main

source for water flowing through the Ombai Strait and Timor Passage.

3.3. The Analysis of Temperature and Salinity Distributions

Consider the aforementioned three major routes of flow of Pacific water through

the Indonesian Seas, which we will identify as the western, northern, and eastern

routes. The western and northern routes begin at the MC port, through which

NPW enters the region before splitting into three branches. The first branch of NPW

takes the western route flowing over the Karakelong and Sangihe Ridges, through

the Sulawesi Sea, and along the Makassar Strait, at the southern part of which it

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 73

Figure 3.1. The locations of vertical sections AGHIJKF, ABCDEF, and LMND are shown by dots. Vertical profiles of temperature and salinity are provided at points identified by letters A, ... , N. Sections, at which normal velocity and salinity are analyzed, are shown by solid lines. Numbers give August transports through these sections in Sverdrups (10 6m3/s) in the direction of arrows. All arrows show transport directed normally to sections; slanting of some arrows is made for convenience of plotting. Values of indices i and j are given on the x- and y-axes. The Mindanao Current ( MC), North Equatorial Counter Current (NECC), New Guinea Coastal Current (NGCC), and Indian Ocean (10) ports are seen where transports enter and leave the model domain ( compare with Figure 2.1(b)).

diverges. Part of this flow exits the region through the Lombok Strait, but the main

portion passes through the Flores Sea to the Banda Sea (see path AGHIJKF in

Figure 3.1). The second branch of NPW takes the northern route flowing through

the Morotai Basin, Maluku Sea, Lifarnatola Passage, Seram Sea and finally into

the Banda Sea, where waters taking the western and northern routes mix (see path

ABCDEF in Figure 3.1). The third branch of NPW leaves the region through the

NECC port. The eastern route begins at the NGCC port, through which SPW enters

the region. Part of this water leaves the region through the NECC port while the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 74

remainder flows through the Halmahera Sea to the Seram Sea where it meets the

northern route of NPW (see path LMND in Figure 3.1). Finally, Banda Sea water

exits the region through the Ombai Strait and Timor Passage.

What factors shape the simulated potential temperature, 9, and salinity, S, dis­

tributions in the Indonesian Seas region? We will show that 9 and S distributions

are strongly influenced by the bottom topography of the region. The main impact

of bottom topography is manifested in the deep layers (greater than 500m depth),

although some influence of the bottom topography on the salinity distribution can be

also traced in the upper 500m. We will disclose the role of some sills and passages in

the formation of 6 and S distributions. In particular, we will come to the conclusion

that 9 and S distributions in the deep layers of the Banda Sea are essentially shaped

by NPW flowing across the Lifarnatola Passage.

It is important to stress that the main structure of the observed distributions

of 9 and S, namely the monotonic decrease of temperature with depth and the

presence of a maximum and minimum in salinity profiles, is satisfactorily reproduced

by the model throughout the entire region. In the upper layer (0 — 500m) the model

and observed profiles are very similar. In the deeper layers temperature deviations

can reach 1 °C while salinity deviations are essentially on the order of several 0.01,

although in some layers, e.g. the layer of salinity minimum at the NGCC port, they

can reach as much as 0.1. Note that 9 and S distributions at the ports are closely

related but not identical to climatology (Conkright et al. 2002) - a result of the

chosen boundary conditions at the open ports (see Chapter 2). The comparison of

simulated and observed data is discussed in detail in Section 5.

Elevations of bottom relief break the Indonesian Seas region down to separate

basins having different 9 and S stratifications. We will analyze the August simulated

data and begin with the western route.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 75

500

1000

1500

2000 Mindanao port(A) E. Sulawesi Sea(G) f 2500 x N. Makassar(H) £ * S. Makassar(l) + Flores(J) 3 3000 O W. Banda(K) 0 S. Banda(F) 3500-

4000-

4500

5000

5500. 8

Figure 3.2. Vertical profiles of potential temperature along the western route for August at points A, G, H, /, J, K, and F (a) (above) and the distribution of potential temperature on vertical section AGHIJKF (b) (shown in Figure 3.1). The bottom topography along this section is shown. Note the break of vertical scale at 300m. Locations of points A, G, H 1 /, J, K, and F are indicated along the top of the figure. Distance from point A is shown on the x-axis. Geographical names and locations are shown (see also Figure 2.1(b)).

3.3.1. Western Route

Consider first the transformation of temperature profiles between points A, G, H,

I, J, K, and F (Figure 2a). In the upper 50m layer the profiles differ slightly due to

basically the surface boundary condition. Then in the layer of strong thermocline (50-

300m) and even deeper (up to 500m) all profiles are similar. Noticeable differences

appear at ~ 1000m. For depths greater than 1000m G-profile has substantially higher

temperature than initial /I-profile; G-. H-. and I- profiles practically coincide; and

J-profile (the Flores Sea) slightly diverges from /-profile. Below 2000m K- profile

(west Banda Sea) diverges from J-profile, but K- and F- profiles within the Banda

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 76

AG HI J K F

100

200

300

1000

J 7 T ,4.75

,4 5 4.5

® 2000 Dewakang 4 25

3000 N Makassar Sangihe strait ridge

4000 Flores sea 175/ Karakelong ridge and basin Sulawesi Banda sea 5000 Pacific Ocean o 500 1000 1500 2000 2500 3000 distance(km)

Figure 3.2(b).

Sea coincide. Temperature in the deep Flores Sea is higher than in the Sulawesi Sea

and Makassar Strait. Notice that bottom water in the Banda Sea is substantially

colder than that in the Flores Sea (compare J-profile with K- and F-profiles). The

above mentioned features are clearly seen in the vertical section of temperature taken

along AGHIJKF (Figure 3.2b). In the upper 500m isotherms are not interrupted

by the bottom topography. So the thickness of the thermocline does not change sub­

stantially although isotherms appear elevated over the Sangihe Ridge and Dewakang

Sill.

Examine the divergence of A- and G-profiles below 1000m. Note that the section

goes through a passage in the Sangihe Ridge with a depth of ~ 1200m. Water

above the depth of the passage can easily overflow the ridge and enter the adjacent

basin. Below the depth of the passage, it is likely that fluid particles to the east

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 77

of the Sangihe Ridge also overflow the ridge due to the strong upwelling seen there

(see further the discussion of Figure 3.8 ). Due to the overflow all isotherms with

6 < 4°C are blocked by the ridge while isotherms with 6 > 4°C pass over it. It

is convenient to identify the 4°C isotherm as a threshold isotherm. Thus, a quite

different 6 stratification is shaped downstream of the ridge relative to upstream.

What characteristics of the upstream flow and underwater topography determine

such a stratification? There is a vast literature on overflow dynamics (see, e.g., Pratt

& Lundberg 1991; Killworth 1994; Pratt & Chechelnitsky 1997; Nielsen et al. 2004

and references therein), but we will not attempt to answer this question in detail.

We will simply demonstrate different types of overflows in the Indonesian Seas region

and provide some estimate of advection (of, say, salinity) into the basin considered.

Note that; the approach based on revealing blocked and passed isolines is often used

in the interpretation of observed hydrographic data (see, e.g., Gordon et al. 2003(a)

and references therein).

The slight distinction between I- and J-profiles is due to the topographic rise

between the Sulawesi Sea and the Makassar Strait and Flores-Banda Ridge. The

topographic rise blocks isotherms with 6 < 4.5°C' while, as we show later, some cold

water flows around and over the Flores-Banda Ridge into the Flores Sea from the

Banda Sea. Following our approach, we cannot explain the departure of J- and K-

profiles by assuming that the overflow of Flores water across the ridge between the

Flores and Banda Seas shapes the distribution of temperature in the Banda Sea. The

only possible explanation is that colder bottom water in the Banda Sea is formed

along the northern route ABCDEF.

We now turn to the analysis of salinity profiles (Figure 3.3a). Because these

profiles contain more structure than their temperature counterparts, they will provide

us with some additional information about the dynamics. The distinguishing feature

of kl-profile, characteristic of NPW, is the presence of a maximum (34.74) at a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 78

500

1000

1500

Mindanao port(A) 2000 E. Sulawesi Sea(G) x N. Makassar(H) 1 2500 • S. Makassar(l) £ + Flores(J) -g 3000 O W. Banda(K) 0 S. Banda(F) 3500

4000

4500

5000

34.1 34.2 34.3 34.4 34.5 34.6 34.7 34.833.9 S

Figure 3.3. Vertical profiles of salinity along the western route for August at points A, G, H, I, J, K, and F (a) (above) and the distribution of salinity on vertical section AGHIJKF (b) shown in Figure 3.1. Further explanations are similar to those in the caption to Figure 3.2.

depth of ~ 150m and a minimum (34.46) at a depth of ~ 350m. The structure of

G-, H-, /-, J-, and A'-profiles in the upper ocean is basically the same. They have

maxima and minima at approximately the same depths, however the magnitudes of

these extrema differ substantially. The magnitudes of maxima markedly decrease

(when one moves from A-profile to /7-profilc). while the magnitudes of minima grad­

ually increase. It is quite possible that this is due to the effect of vertical diffusion.

F-profile has practically no extrema. A more detailed picture is given by the vertical

section of salinity along AGHIJKF (Figure 3.3b). The high salinity signature of

NPW extends as far as the Dewakang Sill but is interrupted at the Karakelong and

Sangihe Ridges probably due to the upwelling that occurs near the Sangihe Ridge.

The strong lowering of isohaline 34.5 over the Dewakang Sill causes the signature of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 79

AG H I J K F

34-4

100 165

200 34.55

300

1000

34 58

® 2000 Dewakang

3000 N. Makassar

4000

Flores Karakelong Sulawesi ridge and basin Banda 5000 Pacific Ocean o 500 1000 1500 2000 2500 3000 distance(km)

Figure 3.3(b).

high salinity to end there.

In general, Figures 3.2 and 3.3 demonstrate a gradual evolution of NPW in the

upper 500m along the western route. No intrusion of water with substantially dif­

ferent temperature and salinity characteristics can be identified. Because the main

part of the Indonesian Throughflow passes through the Makassar Strait we conclude

that basically it is NPW that shapes the temperature and salinity structure of the

upper part of the Indonesian Throughflow, as was stated by Wyrtki (1961); Gordon

(1986); and Gordon & Fine (1996).

We will analyze now in detail salinity G-, H-, J-, K-, and F- profiles below

500m. First, G-, H-, /-, J-, K-, and F-profiles appear to be saltier than Wprofile

in the depth range 500 — 1100m. Then we observe that G'-profile becomes less salty

than Wprofile below 1500m. This is also clearly seen in Figure 3.3b. There is no

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. j=214-226

Figure 3.4. Isotachs of normal velocity (color) and isohalines (solid lines) at the Sangihe section (*=156; the values of j are indicated on the x-axis). Velocities are in m/s. For reference see Figure 3.1. The bottom topography is shown. The western route section (Figures 3.2(b), 3.3(b)) passes through this section at j=223.

doubt that the deviation is caused by the blocking effect of the Sangihe Ridge,

although such a separation begins at a deeper level than for temperature profiles.

The determination of threshold isohalines and levels, at which the corresponding

salinity profiles diverge, is helpful in identifying the impact of topographic barriers

dividing the region into separate basins. Yet these characteristics do not allow us to

estimate a range of salinities advected into a basin. To explain the role of topographic

dividers in shaping distinct stratifications in different basins it is more informative

to consider the distribution of normal velocity and salinity on sections across the

flow. As seen in Figure 3.4, the deep flow at the passage through the Sangihe Ridge

is directed westward. The highest salinity found near the bottom of this passage is

S ~ 34.59. The next barrier is the Dewakang Sill across which J-profile (Flores Sea)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 81

Figure 3.5. The same as Figure 3.4 but for the Dewakang section (j = 138; the values of i are indicated on the x-axis). The vertical line shows the location of the section through the Dewakang sill (j=50) presented in Figure 8 . The western route section (Figures 3.2(b), 3.3(b)) passes through this section at i=57.

becomes slightly more salty than /-profile below ~ 600m. The salinity section in

Figure 3.3(b) as well as the temperature section (Figure 3.2(b)) do not distinctly

reveal the effect of the Dewakang Sill. The section across the Dewakang Sill (Figure

3.5) appears more informative. It shows that the southward flow carries water of

S < 34.54, so the blocking effect of the sill is quite obvious. Note that salinity in

the deep Flores Sea is higher than 34.58 which is due to the penetration of saltier

water from the Banda Sea, as was indicated by Gordon et al. (1994); Top et al.

(1997); and Gordon et al. (2003(a)). This was also suggested by the consideration

of temperature profiles but not in such a pronounced way. Indeed, the cross-section

through the Flores-Banda ridge (Figure 3.6) shows some deep westward flow in the

south (j=103-105) and north (j=112-135) carrying water of salinity S ~ 34.58 and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 82

i

Figure 3.6. The same as Figure 3.4 but for the Flores-Banda section (/= 78; values of j are indicated on the x-axis). The western route section (Figures 3.2(b), 3.3(b)) passes through this section at j= l 10.

higher. As in the case of #, the distribution of salinity in the deep layers of the Banda

Sea demonstrates that saltier waters, relative to the Flores Sea, can come from the

northern Banda Sea only. Finally, we provide Figure 3.7 to show that the entire

water column moves over the Dewakang Sill from the north to the south without any

return flows.

It is known that the calculated ’vertical’ velocity wpom used in the Princeton

Ocean model (POM) is very noisy. The results of calculation of real vertical velocity

w will probably be even noisier. Nevertheless, we will try to extract some information

from the calculated wpom (Figure 3.8). First, it follows from formula (D.3) in the

Appendix that wpom is the difference of the normal components of velocities of the

fluid particle at a point M of the a-surface and of the point M of that surface itself.

The factor N in (D.3) is very close to unity (see (D.2)). Second, near the bottom

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 83

100

2 0 0 -

300

•£■400

■8 500

600 0.1 m/s

700

800

135 136 137 138 139 140 141 142

Figure 3.7. v-velocities in m/s at the section through the Dewakang sill (*= 50; values of j are indicated on the x-axis). See Figure 3.5 for the location of this section. The bottom topography is shown. The numbers on the x-axis (j) should be reversed with the largest on the left and the least on the right.

cr-surfaces do not practically move. Taking into account that the direction of the

normal vector n a (see (D.2)) is close to the positive vertical direction, we conclude

that both velocities wpom and w almost coincide near the bottom. Now consider

Figure 3.8, restricting ourselves to the analysis of signs of the bottom vertical velocity

only. We see that there is indeed upwelling over the Sangihe Ridge and downwelling

over the southern part of the Dewakang Sill. Downwelling also occurs over the eastern

side of the Flores-Banda ridge. Thus, NPW water first overflows the Sangihe Ridge,

and then the Dewakang Sill (Figure 3.7) and the Flores/Banda ridge. Note that high

values of wpom are caused by very steep bottom topography. As we see from Figure

3.8 the slope of the bottom topography is on the order of 10~2.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 84

distance(km)

Figure 3.8. The distribution of the POM ’vertical’ velocity wpom on vertical section AGHIJKF (see Figure 3.1). Velocities are in m/s. The results are noisy but signs of wpom are provided at some points. Bottom topography along this section is shown. Distance from point A is shown on the x-axis. See Figure 3.3(b) for geographical names.

3.3.2. Northern Route

Consider now the formation of 6 and S in the Seram Sea (point D) beginning

with part ABCD of the northern route, see Figure 3.1. Compare first temperatures

of A- and B-profiles (Figure 3.9a). These profiles basically coincide down to a depth

of 1500m and slightly diverge in deeper layers. Figure 3.9b clearly shows that the

threshold isotherm caused by the steep rise of the bottom topography is 3°C. The

water of C-profile is warmer than water of 5-profile below 2000m. This can be

explained by the blocking effect of the topographic rise between these profiles. C-

and D-profiles are separated by the Lifamatola Sill. We see that C- and D-profiles

diverge below 1300m. Figure 3.9b shows that isotherms between 5 — 4 °C appear to

be inclined over the Lifamatola Sill, leading to the separation of profiles at a depth

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 85

500

1000

1500

2000

£ 2500 Mindanao port (A) N. Morotai basin (B) 3000 • S. Maluku sea (C) + Seram sea (D) 3500 x N. Banda sea (E) O S. Banda sea (F) 4000

4500

5000, e

Figure 3.9. Vertical profiles of potential temperature along the northern route for August at points A, B, C, D, E, and F (a) (above) and the distribution of potential temperature on vertical section ABCDEF (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2.

less than that of the sill. Note that we did not see such an effect across the Dewakang

Sill. As seen in Figure 3.9b, the threshold isotherm across the Lifamatola Sill is

3.75°C.

The salinity profiles have a more complicated structure in the upper layer (Figure

3.10a). The same maximum that was discussed in the analysis of the western route

is seen in all profiles at depths between 100 — 200m, and in C- and D-profiles the

magnitude of this maximum is even greater, probably due to the penetration of

saltier water through the Serarn-Obi, Obi-Halmahera, and Lifamatola Passages. But,

as shown below, this effect is not properly described by the model. The minimum

is rapidly destroyed owing to the action of vertical diffusion. In Figure 3.10b, the

elongated patch of high salinity is interrupted over the topographic rise between A-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 86

A B C D E F

100

200

300

1000

425.

3000

Maluku Morotai basin 4000

Seram Pacific 5000 S. Banda N. Banda

o 200 400 600 800 1000 1200 1400 1800 2000 distance(km)

Figure 3.9(b).

and 5-profiles (due to upwelling, see Figure 3.11). The comparison of A- and B-

profiles, B- and C-profiles and C- and 5-profiles below 500m does not give any

substantial new information about the dynamical role of the topographic rise and

barriers between A- and 5-profiles, when compared with the analysis of temperature

profiles. Note that C- and 5-profiles diverge below 1300m. The effect of the Lifam­

atola Sill works below 1300m as in the case of temperature (we observe here some

inclination of isohalines as well). The topographic rise between C- and 5-profiles

(Lifamatola Sill) blocks isohalines with S > 34.62. Figure 3.13 shows very clearly

that the highest salinity advected southward across the sill is 34.61. We will show

that this overflow shapes the salinity in the deep layers of the Seram Sea. The veloc­

ity structure of the water overflowing across the Lifamatola Sill is seen in Figure 3.14.

It is interesting to indicate that seemingly not all water can overflow this sill. We

see a near bottom recirculation on the northern side of the Lifamatola Sill implying

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 87

O O fl1' l"1 Xf 500

1000

1500

2000

f 2500

u 3000

3500 Mindanao port (A) N. Morotai Basin (B) * S. Maluku sea (C) 4000 + Seram sea (D) x N. Banda sea (E) 4500 O S. Banda sea (F)

5000

34.234.1 34.3 34.534.4 34.6 34.7 S

Figure 3.10. Vertical profiles of salinity along the northern route for August at points A, B, C, D, E, and F (a) (above) and the distribution of salinity on vertical section ABCDEF (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2.

that probably there is not a complete flushing of the basin located to the north of

the sill.

3.3.3. Eastern Route

Consider now the eastern route, LMND , see Figure 3.1. L- and M-profiles of

temperature are very similar down to a depth of 3000m (Figure 3.15a). A-profile

(Halmahera Sea) diverges from M-profile below 500m due to the northern Halmahera

Sill of depth 600m. Below the sill, the temperature of iV-profile is higher than that

of M-profile. D-profile (Seram Sea) diverges from A-profile at the depth of the

southern Halmahera Sill (600m). But the temperature of D-profile is colder than

that of A-profile (Halmahera Sea). It seems that the deep Halmahera basin has its

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 88

A B C D E F

34.4 34.2

100

200

I.55 300

1000

34.6

2000 34.62 J4.51

3000

Maluku Morotai basin 4000 1.66

Seram S. Banda 5000 Pacific

o 200 600400 800 1000 1200 1400 1600 1800 2000 distanee(km)

Figure 3.10(b).

own dynamics, being isolated from the Pacific Ocean to the north and the Seram

Sea to the south by the northern and southern sills (of about the same depth). This

is why both the northern and southern sills impact the flow towards the Seram Sea

jointly as if the Halmahera Sea does not exist. These features are clearly seen in

Figure 3.15b.

The analysis of salinity profiles gives some additional information about dynam­

ical processes in this area (Figure 3.16a). First, we see a strong salinity maximum

of 35.50 in L-profile located at ~ 100 — 150m with a weak minimum below 1000m.

This structure is characteristic of SPW. M- and D-profiles also have maxima at the

same depth but with markedly reduced magnitudes. As discussed previously, this is

probably explained by vertical diffusion. It is interesting to note two strong lowerings

of isohalines over the northern and southern Halmahera Sills (Figure 3.16b). We

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 200 400 600 800 1000 1200 1400 1600 1800 2000 distance(km)

Figure 3.11. The same as in Figure 3.8 but for the vertical section ABCDEF (see Figure 3.1).

observed a similar lowering over the Dewakang sill. Note also that we have not seen

anything similar in the temperature section (Figure 3.15b).

A weaker salinity maximum is seen in the Halmahera Sea compared to L- and

M-profiles. It is interesting to note that the magnitude of the salinity maximum in

the Seram Sea is slightly higher than that in the Halmahera Sea, but below 500m

the salinity in the Halmahera Sea becomes higher than the Seram Sea. This again

enforces the idea of an isolated Halmahera basin. Below 1500m, the Seram Sea

water (D-profile) is fresher than water to the north of the northern sill (M-profile).

A similar anomaly was observed in the analysis of temperature profiles.

Next we analyze the structure of currents and salt transports through the Halma-

hera Sea in more detail. Consider Figure 3.17. There are three main branches of the

flow: the near bottom southward current advects salinity 34.66; the near surface

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 90

135 140 145 150 155 160 i

Figure 3.12. The same as in Figure 3.4 but for the North Maluku section (j=187; values of i are indicated on the x-axis). The northern route section (Figures 3.9(b), 3.10(b)) passes through this section at i—151.

southward current (in the western part of the section) advects salinities 34.66 and

less; and the northward current (in the eastern part of the section) advects higher

salinities (up to 34.90). Figure 3.18 shows a section across the Halmahera Sea. To

the west and to the east of the section, the continuation of southward and northward

currents identified in Figure 3.17 can be observed. In the central deep part of the

sea we see anticyclonic circulation that supports the idea of the isolated Halmahera

Sea. Consider now the section across the southern Halmahera Sill (Figure 3.19). We

see the previously identified southward and northward currents in the western and

eastern parts of this section. It appears that the overflow from the Halmahera Sea

has a rather complicated character. According to Figure 3.20, water sinks on the

southern side of both sills but downwelling and upwelling are intermittent on the

northern side of the southern sill. Finally the Obi-Seram section (Figure 3.21)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 91

distance(km)

Figure 3.13. The same as in Figure 3.4 but for the Lifamatola section (j = 150; values of i are indicated on the x-axis). The vertical line gives the location of the section through the Lifamatola passage (i=139) in Figure 3.14. The northern route section (Figures 3.9(b), 3.10(b)) passes through this section at i=137.

shows a more organized velocity pattern.

In the upper 500m water of rather high salinity (up to 34.72) is advected into

the Seram Sea. A major part of this flow enters the Banda Sea through the Buru-

Mangole passage (see Figure 3.22) with a maximum salinity of 34.75 while most

of the remainder enters the Maluku Sea as a weaker northward current across the

Lifamatola Passage also with salinity maximum 34.75 (see Figure 3.13). The analysis

of isohalines and isotachs of normal velocity at the New Guinea-Seram, Buru-Seram,

and Flalmahera-Obi sections (not shown) also suggests the existence of a secondary

circulation pattern. A slow northward motion through the New Guinea-Seram pas­

sage carries water of salinity 34.50 - 34.70. This current continues northward before

making a loop and joining the main flow near the deep Halmahera basin and the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 92

500

1000

£ 1500

2000

0.1 m/s 2500

146 148 150 152 154 156

Figure 3.14. v-velocities in m /s at the section through the Lifamatola passage (i= 139: values of j are indicated on the x-axis). See Figure 3.13 for section location. The bottom topography is shown. The numbers on the ./-axis (j) should be reversed with the largest on the left and the least on the right.

south Halmahera Sill and returning towards the Seram Sea. One branch of this

current leaves the area between Obi and Halmahera islands (carrying salinity 34.60 —

34.65) while the remainder enters the Seram Sea. We mention also a weak southward

current between Buru and Seram Islands having salinity 34.50 — 34.65. We did not

find any noticeable transport of SPW entering the Maluku Sea from the north of

Halmahera Island.

The appearance of high salinity water in the Obi-Seram section, in the Seram Sea

and to the north of the Lifamatola Passage is observed and usually interpreted as a

signature of SPW (see, e.g., Aken et al. 1988; Gordon and Fine 1996). Unfortunately,

the results of our model appear to be inconclusive on this point. This is connected

with some specifics of our model: the nudging of S to the climatology used in our

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 93

------500

1000

2000 NGCC port (L) N.E. of Halmahera Sea (M) • Halmahera Sea (N) 2500 * Seram (D)______

3000

3500, 9

Figure 3.15. Vertical profiles of potential temperature along the eastern route for August at points L, M, N, and D (a) (above) and distribution of potential temperature on vertical section LMND (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2.

simulations. This climatology was based on the representation of realistic climatology

by a certain number of Fourier harmonics (see Chapter 2) that resulted in some

increase of climatological values of S in the upper 0 — 500m in this area. This is

the reason why we observed high simulated salinities in the Obi-Seram section, in

the Seram Sea and to the north of the Lifamatola Passage (see, e.g., profile C in the

South Maluku Sea, Fig 3.11a), although high values of salinity are not seen on the

section through the southern Halmahera Sill. Nevertheless, based on the direction of currents in the upper layers of the Obi-Seram section and Lifamatola Passage it

is plausible that some SPW can penetrate to the Seram and Maluku Seas.

But the correspondence between the two climatologies below 500m is satisfactory.

Below 500m water with S = 34.61 enters the Seram Sea largely through the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 94

100

200

300

■o

2000 4-25 —_ ------F-'-y3.25 S. Halmahera

3000 i £ - Halmahera

4000, 200 400 600 800 1000 1200 distance(km)

Figure 3.15(b).

Lifamatola Passage (see Figure 3.13). In the Obi-Seram section (Figure 3.21) a

current, with salinity 34.61, is directed eastward (out of the Seram Sea). Furthermore,

this water flows northward through the south Halmahera Sill and eastward through

the New Guinea-Seram passage. There is a weak northward current between Buru

and Seram Islands with salinity 34.61, which essentially plays no role in the salinity

budget of the Seram Sea. The net flow between Mangole and Buru Islands is out of

the Seram Sea and into the Banda Sea. The flow of bottom water from the Seram

Sea into the Banda Sea is located to the south of the Mangole-Buru section and has

a salinity of 34.61. There is, however, a weak gyre in the northern part of the section

which clearly is of no importance to the salinity budget of the Seram Sea. Thus, we

conclude that the salinity in the deep part of the Seram Sea is formed by the flow of

NPW across the Lifamatola Passage. This conclusion agrees with the presently

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 95

x« 500

1000

NGCC port (L) N.E. of Halmahera Sea (M) 2000 • Halmahera Sea (N) * Seram (D)______

2500-

3000

3500 34.2 34.4 34.6 34.8 35.2 35.4 S

Figure 3.16. Vertical profiles of salinity along the eastern route for August at points L, M, N, and D (a) (above) and the distribution of salinity on vertical section LMND (b) (see Figure 3.1). Further explanations are similar to those in the caption to Figure 3.2.

accepted view (see, Aken et al. 1988, 1991; Gordon et al. 2003(a)). It is the advection

of water of salinity 34.61 through the passage between Mangole and Buru that forms

the salinity of the deep Banda Sea (see also Figure 3.3b).

We now complete the discussion of the northern route. Below 2700m A-profile

(north Banda Sea) is colder than D-profile water (Seram Sea). We cannot explain

this following the approach based on the consideration of overflows. In Figure 3.11

we saw that in the deep north Banda Sea there is upwelling that is probably balanced

by vertical diffusion. Due to this balance, the temperature of this profile decreases to

the bottom. It is possible that bottom temperature in the Banda Sea is established

by this effect. The salinities of E- and D-profiles are very similar.

Finally, we consider velocity and salinity distributions across the Lombok and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 96

L M N D

J4.4 c 100

200 14.66

300

1.64, f 1000

W.63

2000

S. Halmahera

3000

Pacific Seram

200 400 600 800 1000 1200 distance(km)

Figure 3.16(b).

Ombai Straits and Timor Passage (Figures 3.23 - 3.25). Basically, the normal velocity

at all exit passages is unidirectional (outflow) although very weak opposing currents

are seen in the Ombai Strait (between 500 and 1000m) and Timor Passage (near

the bottom, deeper than 400m). In all three cases salinity increases with depth and

no internal maxima or minima are observed. The maximum salinity in the Lombok

Strait is 34.55, in the Ombai Strait 34.59 and in the Timor Passage 34.61.

3.4. Seasonal Variation

There are no substantial structural changes of potential temperature, 0, and

salinity, S, distributions between seasons, though values of some parameters of these

distributions (e.g., magnitudes of maxima and minima) can change. The comparison

of corresponding profiles shows that seasonal variability is greatest near the surface.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 97

Figure 3.17. The same as in Figure 3.4 but for the northern Halmahera sill {j~ 166; values of i are indicated on the z-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=171.

Surface temperature changes seasonally by up to ~ 2°C in the northernmost and

southernmost regions but these changes are less pronounced away from these regions.

Seasonal changes in surface salinity are generally greatest near land, particularly in

the Makassar Strait region and around the Halmahera and Seram Seas. In the

thermocline, temperature changes on the order of ~ 0.5°C' occur while isotherms

can be displaced vertically by several tens of meters. Salinity extrema generally

exist throughout the year but their values can change somewhat and they can be

vertically displaced. However, the salinity maximum seen in the Seram and Maluku

Seas in August is much weaker in February and disappears in May and November.

Strong raising and lowering of isohalines, generally associated with topography, is

usually seen throughout the year. Below the thermocline, blocking of some isotherms

and isohalines is observed throughout the year. Though changes of temperature and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 98

154 156 158 160 162 164 166 168 170 i

Figure 3.18. The same as in Figure 3.4 but for the Halmahera sea section (j=157; values of i are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=164.

salinity values are very small, isotherms and isohalines can be displaced by several

tens of meters. Velocity values across the major deep overflows vary little by season

but there is some change in the magnitude and extent of deep downwelling and

upwelling cells that may explain the displacement of deep isotherms and isohalines.

Thus, our conclusions regarding the formation of deep salinities in the Seram and

Banda Seas based on August data remain valid throughout the year. The following

comparisons refer to results in February and all changes are considered relative to

August unless otherwise stated.

3.4-1- Western Route

The structure of 6 profiles along the western route does not change seasonally.

Temperature change is greatest in the upper mixed layer ranging from an increase

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 99

Figure 3.19. The same as in Figure 3.4 but for the southern Halmahera sill (j = 149; values of i are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at i=156.

of ~ 1 °C at the MC port (A-profile) to a decrease of ~ 2°C in the south Banda Sea

(F-profile). In the thermocline, there is very little seasonal variability at the MC

port and in the Banda Sea. Temperatures decrease by ~ 0.1°C in the east Sulawesi

Sea (G-profile) and by ~ 0.5°C in the Flores Sea (J-profile) and increase by ~ 0.1°G

in the west Banda Sea (F-profile). A comparison of temperature sections shows

that isotherms within the thermocline are raised by ~ 10 — 20m from the MC port

(A-profile) to the Sangihe Ridge. Between the Sangihe Ridge and the Dewakang Sill

the isotherms are lowered by ~ 10 — 20m in the upper thermocline while they are

raised by ~ 10 — 20 m in the lower thermocline. In the Flores Sea and west Banda

Sea isotherms within the main thermocline are lowered by ~ 20m. In deep water,

some notable vertical seasonal displacement of isotherms occurs. In the Sulawesi Sea

isotherms are raised by ~ 50m down to a depth of ~ 1750m. Below this depth

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 100

distance(km)

Figure 3.20. The same as in Figure 3.8 but for the vertical section LMND (see Figure 3.1). Distances from point L are shown on the ;/:-axis.

isotherms are displaced upwards by ~ 100m except in the region of the Sangihe

Ridge, and the conjunction of the Sulawesi Sea with the Makassar Strait, where

displacement is markedly reduced.

For salinity profiles, there is some seasonal change in surface and extrema values

but essentially no change in the deeper structure. In the mixed layer, the maximum

seasonal salinity change occurs in the south Makassar Strait where it increases by

~ 1.5 while the Flores Sea freshens by ~ 0.7. Elsewhere, seasonal changes are not

as pronounced. Surface values increase at the MC port (A-profile) and the south

Banda Sea (F-profile), while they decrease in the east Sulawesi Sea (G-profile) and

the north Makassar Strait (//"-profile). Seasonal changes seen near the surface and in

the mixed layer in the south Makassar region are the direct result of monsoon wind

forcing, see, e.g., Gordon et al. (2003(b)). In February, northwest monsoon winds

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 101

i

Figure 3.21. The same as in Figure 3.4 but for the section between Obi and Seram islands (£=148; values of j are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at j=144.

move near surface low salinity water into the south Makassar region. In August,

the southeast monsoon winds force freshwater out of the south Makassar Strait.

The salinity maximum at A-profile is located at 100m (125m in August) while the

minimum is located at 400m (300m in August). The salinity minimum of G'-profile

is also situated at 400m (300m in August). In the south Makassar Strait (I-profile),

a very weak Smax at 200m replaces the strong maximum seen at 100m in August

while a fresher weak minimum at 300m replaces the weak minimum seen at 500m

in August. In the Flores Sea (J-profile), the Smax is weaker and is located at 150m

(100m in August). In the west Banda Sea (.A-profile) the Smax is seen at 150m

(200m in August). Comparing seasonal sections of salinity (figure not shown), we

see a deeper Smax between the Sangihe Ridge and north Makassar. There is no Smax

at the entrance to the Makassar Strait because of the strong upwelling cell there.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 102

i

Figure 3.22. The same as in Figure 3.4 but for the section between Mangole and Buru (i— 128; values of j are indicated on the x-axis). The eastern route section (Figures 3.15(b), 3.16(b)) passes through this section at j=144.

In August, a slight deepening of the Smax is seen in the Sulawesi Sea, while it is

raised strikingly at the Labani Channel. Analysis of vertical velocity along the

western route reveals large seasonal variability in upwelling and downwelling and

it is hypothesized that this is responsible for salinity changes between the Sangihe

Ridge and the Dewakang Sill. An upwelling cell that extends through the water

column at the conjunction of the south-west Sulawesi Sea and northern Makassar

Strait is much stronger in August; a weak upwelling cell at the location of the Smax

interruption in the north Makassar Strait is not present in August; the upwelling cell

in the Labani Channel is much stronger in August. The patch of low salinity water

seen above the Dewakang Sill in August is more extensive in February due, perhaps,

to freshening of the sea surface but also due to the weakness of the upwelling cell.

In August, the winds are directed towards the north along the Makassar Strait and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 103

Figure 3.23. The same as in Figure 3.4 but for the section across the Lombok Strait (j = 110; values of i are indicated on the x-axis).

can generate the upwelling cells at the Dewakang Sill and Labani Channel. In deep

water, isohalines are generally lowered in the eastern Sulawesi Sea, raised in the

western Sulawesi Sea and close to the Makassar Strait but there is little or no dis­

placement close to the Sangihe Ridge.

3.4-2. Northern Route

We now discuss 6 profiles and sections along the northern route. Seasonal tem­

perature change is greatest in the upper mixed layer with Banda Sea water warmer

by 2°C, Seram and Maluku Seas water warmer by ~ 1.0 — 1.5°C, and MC port

and north Morotai water cooler by ~ 1.0° C. In the thermocline, seasonal variability

is greatest in the north Banda Sea where water is warmer by more than 0.6°C and

north Morotai basin where water is cooler by ~ 0.4°C. Changes are smaller in the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 104

i

Figure 3.24. The same as in Figure 3.4 but for the section across the Ombai Strait (i= 90; values of j are indicated on the x-axis).

Seram and south Maluku Seas. Isotherms are raised in the thermocline from the MC

port to the Morotai basin, while they are lowered in the Maluku and Banda Seas. In

deep water, isotherms are displaced upwards adjacent to the topography separating

the Pacific Ocean from the Morotai basin in response to the strengthening of the

upwelling cell there. In the Seram Sea and northern Banda Sea isotherms between

the depth of the Lifamatola Passage and ~ 3000m are displaced downward. This

result is consistent with a large downwelling cell that is both stronger and more

extensive in February. In the south Banda Sea isotherms are displaced upward consistent with stronger upwelling at that time.

Next, we compare salinity profiles and sections along the northern route. A

notable seasonal change in surface salinity is observed. Surface salinity decreases by

~ 0.3 in the Seram Sea and by ~ 0.25 in the south Maluku Sea and north Banda

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 105

I

Figure 3.25. The same as in Figure 3.4 but for the section across the Timor Passage (i= 90; values of j are indicated on the x-axis).

Sea while it increases by ~ 0.1 in the north Morotai basin. The salinity maximum of

the Seram Sea and the south Maluku Sea is much weaker and there is no maximum in

the north Banda Sea. The salinity section shows that there is no Smax in the Banda

Sea, a very weak deeper Smax in the Seram Sea, and a weaker, broader and deeper

Smax in the Maluku Sea that almost disappears between the north Maluku Sea and

Morotai basin. In the northern Banda and south Maluku Seas isohalines at and

below the depth of the Lifamatola Passage (to a depth of ~ 3000m) are displaced

downward, similar to isotherms. In the deep Banda Sea isohalines are displaced

upward, similar to the upward displacement of isotherms there.

3-4-3. Eastern Route

Comparing temperature profiles and sections along the eastern route shows the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 106

upper mixed layer is cooler by ~ 0.4°C at the NGCC port (L-profile) and by ~ 0.3°(7

to the northeast of the Halmahera Sea (M-profile) while it is warmer by ~ 0.2°C in

the Halmahera Sea (/V-profile). In the thermocline, seasonal temperature change is

~ 0.5°C; cooler in the Halmahera Sea and warmer to the north of the Halmahera Sea.

The temperature section shows that in the thermocline, isotherms generally slope

downwards from the Halmahera Sea to the Seram Sea, whereas they slope upwards

in August. Therefore, in February, isotherms are shallower in the Halmahera Sea and

are deeper on the Pacific side of the north Halmahera Sill. A notable temperature

change at depth is the cooling of the deep Halmahera basin by ~ 0.2°C.

In the upper mixed layer, the NGCC port (L-profile) is saltier by ~ 0.3, M-profile

(northeast of the Halmahera Sea) is saltier by ~ 0.5 while Ar-profile (Halmahera Sea)

is fresher by ~ 0.1. The salinity maximum in the Halmahera Sea is located at 150m

(100m in August). Between the south Halmahera Sill and the Seram Sea the Smax is

deeper and much weaker. The area of relatively low salinity at the north Halmahera

Sill is much weaker. In the deeper Halmahera Sea salinity is reduced to 34.63 — 34.64

compared with 34.65 in August.

3-4-4■ Exiting Passages

In the Ombai Strait, surface temperature is greater than 29°C in February but

is less than 27 °C in August. Isotherms are displaced upwards by ~ 25 m above

~ 200m and downwards by approximatley the same amount between ~ 200 — 500m.

Below 500m the isotherms have the same location in February and August but are

displaced downwards in May and upwards in November by ~ 50 - 100m. Surface

salinity varies very little with a maximum of S ~ 34.4 in February and a minimum

of S ~ 34.0 in August. No salinity extrema are seen in the Ombai Strait. Below

the thermocline the 34.58 and 34.59 isohalines are displaced downwards in February

relative to all other seasons.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 107

In the Timor Passage, surface temperature is greater than 28 °C in February but

is less than 27 °C in August. The thermocline is displaced downwards by ~ 25 m. In

the lower thermocline (below ~ 500m) the isotherms have a greater downward slope

toward the south. Surface salinity exceeds 34.8 in February, has a value of ~ 34.0 in

May and ~ 34.5 in August and November. No salinity extrema are seen in August

or in February. However, there is a weak localized Smax ~ 34.64 at ~ 125m in May

and Smax ~ 34.65 at ~ 100m in November. Below ~ 1000m water of salinity 34.61

is observed but a patch of slightly saltier (S ~ 34.62) water is seen in May.

In the Lombok Strait, surface temperature is greater than 28°(7 but is less than

27°C in August. Below the mixed layer, the isotherms are displaced downward by

~ 20m. The surface salinity is S ~ 34.0 in February and August, slightly less in May

and slightly more in November. No salinity extrema are seen for the whole year.

3.5. Comparison with Observations

In general, the developed model correctly reproduces the basic features of ob­

served potential temperature and salinity distributions. We consider now in detail

results of the comparison of simulated and observed data.

3.5.1. Heat and Salt Transports

Stammer et al. (2003) using the Massachusets Institute of Technology (MIT)

General Circulation Model (GCM) fully constrained by the World Ocean Circula­

tion Experiment (WOCE) data, found mean heat transport of about 1.1 PW from

the Pacific Ocean into the Indian Ocean. Schneider & Barnett (1997) using a cou­

pled ocean-atmosphere model found that heat transport towards the Indian ocean is

maximal in July (1APW) and minimal in February (0.1 PW). Ganachaud & Wunsch

(2000) applying an inverse model to hydrographic and current meter data and using

climatological wind fields and some biogeochemical balances calculated a heat flux

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 108

of 1APW entering the Indian Ocean. Schiller et, al. (1998) using an Ocean Global

Circulation Model (OGCM) with enhanced tropical resolution found a mean heat

transport of 1.15PIE across the section between western Australia and Java towards

the Indian Ocean with a maximum of 1.86PIE in August and a minimum of 0.58PIE

in March. Hirst & Godfrey (1993) applying the Bryan-Cox Ocean General Circula­

tion Model found a heat transport of 0.62PIE out of the Pacific Ocean toward the

Indian Ocean. Lee et al. (2002) calculated a heat flux of approximately 0,6 PIT to

the Indian Ocean. Godfrey (1996) in his review paper suggested that a Throughflow

of 10ST would transport about 0.5PIE of heat into the Indian Ocean. Ffield et al.

(2000) using ARLINDO (ARus Lintas INDOnen) measurements of temperature and

transport from instruments deployed in the Labani Channel estimated a mean heat

flux of 0.50PIF through the Makassar Strait in 1997. They also provided estimates

of 0.63PIE during the La Nina months of December 1996 through February 1997

and of 0.39PIF during the El Nino months of December 1997 to February 1998.

Vranes et al. (2002) calculated a heat transport through the Makassar Strait for the

period December 1996 - July 1998 using four different observed temperature profiles.

Their mean values are 0.72PIT, 0.58PIT, 0.34PIF, and 0.53PJF respectively. Gordon

& McClean (1999) analyzed the regional circulation of the Indonesian Seas provided

by the Los Alamos National Laboratory (LANL) Parallel Ocean Program (POP)

forced by ECMWF (the European Center for Medium Weather Forecasting) wind

stresses for the period 1985 through 1995. They found a heat transport towards the

Indian Ocean of 0.79PIE in August 1993 and 0.06PILr in February 1994 and a salt

transport of 2.51 x 10 skgs~1 in August 1993 and 3.0 x 10 7kgs~'L in February 1994.

Piola and Gordon (1984) applying simple box models suggested a total salt flux from

the Pacific Ocean to Indian Ocean of 4.71 x lO ^ps"1.

3.5.2. Western Route

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 109

We begin with the western route. Essentially, there is satisfactory agreement

between the simulated and observed (Gordon et al. (2003a)) estimates of the effective

depths of topographic features (the effective depth is the depth of the threshold isoline

upstream of the obstacle). There is a good fit of observed to simulated features such

as the location of the salinity maximum of S ~ 34.75 and minimum of S ~ 34.48

at ~ 125m and ~ 300m respectively and the lower boundary of the thermocline

at ~ 250m in the north Makassar Strait (//-profile); the location of the salinity

maximum of S ~ 34.64 at ~ 100m and minimum of S ~ 34.51 at ~ 300m, and the

lower boundary of the thermocline at ~ 300m in the Flores Sea (see Gordon 2005;

Figure 4). However, simulated deep temperatures of 9 ~ 4.2°C in the East Sulawesi

Sea (G'-profile) and 9 ~ 2.0°C' in the MG port region (/I-pro file) are slightly higher

than observed; in contrast, the corresponding simulated salinities of S ~ 34.58 and

5 ~ 34.65 are slightly lower than observed (see Gordon et al. 2003(a); Figure 2).

In shallower waters in the Makassar Strait and the Flores Sea, the model results

are very similar to the data of Ilahude & Gordon (1996; Figure 3). For salinity, the

simulated results agree with data at depth provided by Gordon et al. (2003a; Figure

5(b)), where the deep Makassar Strait and southwest Sulawesi Sea are fresher than

the Flores Sea which in turn is fresher than the western Banda Sea. In shallower

waters, the position of isohalines characterizing the salinity maximum and minimum

along the Makassar Strait are satisfactorily reproduced by the model (Gordon et

al.2003a; Figure 5b; Ilahude & Gordon 1996; Figure 3). The same is true for the low

salinity cell over the Dewakang Sill seen in Ilahude & Gordon (1996; Figure 3). The

weaker maximum seen in February in Gordon et al. (2003a; Figure 5b) and Ilahude

6 Gordon (1996; Figure 3) is also reproduced by the model. The relatively weak

salinity maximum and minimum in the Flores Sea observed by Gordon et al. (1994;

Figure 4) and the absence of salinity extrema in the western Banda Sea (Gordon et

al. 2003a, Figure 5b; Gordon et al. 1994, Figure 4; Ilahude & Gordon 1996, Figure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 110

4) are adequately displayed by the model. We did not find appropriate observations

to check the validity of the lack of a salinity maximum near the Karakelong and

Sangihe Ridges.

3.5.3. Northern Route

Along the northern route, the model profile in the south Maluku Sea (G'-profile)

gives salinity maximum of S ~ 34.77 at ~ 150m, in good agreement with that seen

in Gordon (2005; Figure 4). The same is true for the model salinity minimum of

S ~ 34.56 at ~ 800m and model salinity of S ~ 34.62 below ~ 1600m (compare

with Gordon et al. 2003a; Figure 2). In the Seram Sea (D-profile), the model

gives a salinity maximum in August of S ~ 34.76 at ~ 150m, in contradiction to

observations (Gordon 2005; Figure 4). But below this maximum the model salinity

increases monotonically with depth to a deep value of S ~ 34.61, in very good

agreement with Gordon et al. (2003a; Figure 2). The salinity maximum seen at

D-profile is much reduced in February and is not apparent in November and May,

in close agreement with observations. In the south Banda Sea (F-profile), the model

shows no salinity extrema in the upper ocean agreeing with observations of Gordon

(2005; Figure 4). The deeper salinity of S ~ 34.61 also agrees with observations

of Gordon et al. (2003a; Figure 3). Comparing temperatures at depth between

the south Maluku Sea (C-profile) and the Seram Sea (D-profile) gives a simulated

effective Lifamatola Sill depth of ~ 1500m., in good agreement with the effective

depth of ~ 1650m found by observations (Gordon et al. 2003a).

Comparing 9- and S- sections along the northern route, we see that the temper­

ature structure in the upper ocean agrees with that of Gordon et al. (2003a; Figure

6a). At depth, the model shows that isotherms slope down over the Lifamatola Pas­

sage into the Seram Sea, and even into the Banda Sea, and that the deep waters of

the south Maluku Sea are much cooler than deep waters of the Seram and Banda

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. I l l

Seas, in accord with Gordon et al. (2003a; Figure 6a) and Aken et al. (1988; Figure

9). For salinity at depth, the model isohalines slope downward across the Lifama­

tola Passage and that is why the deep south Maluku Sea appears saltier than the

Seram and Banda Seas, again in agreement with Gordon et al. (2003a; Figure 6b).

The Banda Sea has almost constant salinity at depth also in accord with Gordon et

al. (2003a; Figure 6b) and Aken et al. (1988; Figure 10). In the upper ocean, the

model shows no salinity extrema in the south Banda Sea, corresponding well with

the data of Gordon et al. (2003a; Figure 6b). In the south Maluku Sea, both model

extrema were observed by Gordon et al. (2003a; Figure 6b), although the model

salinity maximum has a higher value especially in August. In the north Maluku

Sea the model salinity extrema have also been observed by Ilahude & Gordon (1996;

Figure 5), although the model salinity maximum is a little higher than the observed

value. Again, we can say nothing certain about the validity of the strong raising

of isohalines between A- B-profiles (Figure 3.10b) because of a lack of appropriate

observations.

3.5.4- Eastern Route

Along the eastern route, we find an effective sill depth across the northern

Halmahera Sill of ~ 500m, somewhat less than ~ 600m observed by Gordon et

al. (2003a; Figure 2). As a result, the model temperature at depth within the

Halmahera Sea is warmer than observed (see also Cresswell & Luick 2001), while the

simulated temperature difference in M- and A-profiles match the observed difference

below the effective depth of the sill. For the deep Halmahera Sea, the model gives

salinity of S ~ 34.65, more than the observations of Gordon et al. (2003a, Figure 2)

who found S ~ 34.60 (see also Riel 1943). The fact that the model effective depth of

the North Halmahera Sill is less than observed appears to be the main reason for the

relatively warmer and saltier deep water seen in the model in the deep Halmahera

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 112

basin. We did not find appropriate observations to check the validity of massive

isohaline lowering over the northern and southern Halmahera Sills (Figure 3.16b).

For completeness sake we mention here the appearance of high salinity water in

the Obi-Seram section, in the Seram Sea and to the north of the Lifamatola Passage

which is an artifact of the model (see discussion in section 3.3).

3.5.5. Exiting Passages

A very limited number of observations of 9 and S have been made at the exit

passages of the Indonesian seas (Timor Passage, Ombai and Lombok Straits). Fieux

et al.(1996) made CTD and CFC measurements in the Ombai Strait and Timor Pas­

sage during the north-west monsoon in 1992 as part of the Java Australia Dynamics

Experiment (JADE) program. Molcard et al.(1996) measured the throughflow from

2 current meter moorings placed in the Timor Passage from March 1992 - April 1993.

They also made a series of CTD casts in the Timor passage for two days in March

1992. Cresswell et al. (1993) investigated the circulation of the using ship­

board CTD, ADCP and dropsonde measurements from October 1987 to Alarch 1988

with complementary data from satellite tracked drifters, continental shelf moorings

and a Nansen bottle survey during 1976. Molcard et al. (2001) deployed a current

meter mooring in the Ombai Strait in November 1995 for one year. They made

CTD casts at four stations across the Strait in December 1995. Murray et al. (1990)

report on observations made in the Lombok Strait and surrounding regions from

January 1985 until March 1986. They deployed 7 current meter moorings, 5 in the

Lombok Strait and 2 in the west Flores Sea, and made 234 CTD casts in the Lombok

Strait, the adjacent area of the Indian ocean and the west Flores Sea. Sprintall et al.

(2003) studied 5 major exit passages of the between December

1995 and May 1999 by using shallow pressure gauge arrays mounted with salinity

and temperature sensors. They also made underway measurements with ADCP and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 113

CTD packages during deployment of the sensor arrays and three subsequent surveys.

The model results in the Timor Passage agree well with the March 1992 CTD

profiles made by Molcard et al.(1996; Figure 12). They found a surface salinity of

S ~ 34.8 in the upper ~ 50m. As one moves down the water column, the salinity

first decreases (S < 34.5) but then increases monotonically to S ~ 34.61 at ~ 1400m.

The temperature profiles show a surface value of 9 ~ 30oC with the 10°C isotherm

located at a depth of ~ 400m. A closer comparison between CTD profiles of Fieux

et al. (1996; Figure 6) and model results between 1000 — 1500m shows that the

salinities match well but the model potential temperature is warmer than observed

by up to 1 °C. Comparing the model results with Cresswell et al. (1993) we find that

the temperature and salinity sections for the upper 300m in October 1987 (Cresswell

et al. 1993; Figure 6a) practically coincide with the model results. Most notably,

the locations and values of the localized salinity extrema are almost identical. For

March 1988, the observed surface temperature is greater than 29°C (Cresswell et al.

(1993; Figure 6b)) is slightly warmer than the model result (8 > 28 °C) but the basic

features of temperature distribution are similar. They also found a surface salinity

of S > 34.8 and salinity minimum of S < 34.45 at depth ~ 100m to the north of

the passage, a feature clearly seen in the model at this time. Sections of 6 and S

down to 200m for August 1976 (Cresswell et al. (1993 Figure 9)) are also in good

agreement with the model results.

The CTD measurements in December in the Ombai Strait by Molcard et al.

(2001; Sect. 3.1, Figure 4) are in good agreement with the model output. In both

cases, the surface potential temperature is greater than 28°(7, the boundaries of

thermocline layers coincide and the 8 °C isotherm is located at 500m. In deeper waters

(below 1500m) the model is slightly warmer than the observed profiles. Salinities are

also in good agreement; surface salinities coincide, but the observations show very

weak extrema of Smax ~ 34.55 at ~ 200m and 5m,n ~ 34.51 at ~ 300m- whereas the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 114

model does not contain any extrema. The simulated salinity increases monotonically

from a surface value of S ~ 34.4 to match the observation below a depth of ~ 300m.

Murray et al. (1990; Figs 7 and 6) show that the salinity extrema of Makassar

water are destroyed upon passing through the Lombok Strait both in June and

September 1985. This feature is well reproduced by the model. Sprintall et al.

(2003) found that surface waters are warmer and fresher in the main 5 exit passages

of the Lesser Sunda Islands from March to May compared to other times of the year.

These results are in agreement with the model output.

3.6. Summary and Conclusions

The formation and distribution of temperature and salinity in the Indonesian

Seas region have been studied with a high-resolution model based on the Princeton

Ocean Model. One of the distinctive properties of the model is an adequate repro­

duction of all major topographic features of the region by the model bottom relief.

It has been stressed in several publications that the formation of temperature and

salinity distributions in the Indonesian Seas, especially in deep water, is strongly

controlled by the bottom topography (see, e.g., Gordon & McClean 1999; Gordon

et al. 2003a and references therein). The model displays satisfactorily the main fea­

tures of potential temperature and salinity distributions in the region. It was shown

that the contribution of the deep parts (below 500m) of the Ombai Strait and Timor

Passage to the total transports of internal energy (heat) and salt mass appears rather

substantial.

To analyze the influence of bottom topography in detail we have considered the

three major routes of flow of Pacific water through the Indonesian Seas, which have

been identified as western, northern, and eastern routes. The western and northern

routes begin at the Mindanao Current (MC) port, through which North Pacific

Water (NPW) enters the region. A branch of NPW, taking the western route, flows

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 115

over the Karakelong and Sangihe Ridges, through the Sulawesi Sea, and along the

Makassar Strait, at the southern part of which it diverges. Part of this flow exits the

region through the Lombok Strait, but the main portion passes through the Flores

Sea to the Banda Sea. Another branch of NPW takes the northern route flowing

through the Morotai Basin, Maluku Sea, Lifamatola Passage, Seram Sea and finally

into the Banda Sea, where waters taking the western and northern routes mix. The

eastern route begins at the New Guinea Coastal Current (NGCC) port, through

which South Pacific Water (SPW) enters the region, flows through the Halmahera

Sea to the Seram Sea, where waters taking the eastern and northern routes mix.

Finally, Banda Sea water exits the Indonesian Seas region through the Ombai Strait

and Timor Passage.

Major topographic features break the region down into separate basins having

different temperature and salinity stratifications. These are the Sangihe Ridge, the

topographic rise between the Sulawesi Sea and Makassar Strait, the Dewakang Sill,

and the ridge between the Flores and Banda Seas along the western route; the topo­

graphic rise between the Morotai Basin and Maluku Sea and the Lifamatola Sill

along the northern route; and the northern and southern Halmahera Sills along the

eastern route. To clearly demonstrate the changes in stratifications we analyzed a set

of temperature and salinity profiles and the corresponding sections along each route.

What are the reasons for these changes? The water spills over major topographic el­

evations, as exemplified by the calculation of vertical velocity over the slopes of these

elevations. The dynamics of the overflow is such that topographic features impede

the advection of low temperature and high salinity from upstream of the topographic elevation to downstream. It seems likely that there is some relation between blocked

ranges of temperature and salinity and characteristics of upstream flow and bottom

elevations. But we did not attempt to seek such a relation and have restricted our­

selves to a description of different types of overflows observed in the Indonesian Seas.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 116

For this purpose we have chosen salinity as a tracer and considered sections above

topographic features or, in other words, at the boundaries between basins. Isohalines

and isotachs of normal velocity at these sections have been presented and discussed.

Our assumption was that the highest salinity in the downstream basin, below the

depth of the topographic feature, should be advected from the neighboring basins.

In this way, we have disclosed, for example, the blocking effect of the Dewakang Sill.

This effect was hardly seen from a traditional consideration of the threshold isohaline

or isotherm. Following such an approach we were able to prove that the difference in

temperature and salinity stratifications is really caused by the impact of the bottom

topography.

The salinity stratification of NPW and SPW entering the Indonesian Seas region

are quite different. NPW is characterized by a salinity maximum of ~ 34.74 at

a depth of ~ 150m, a deep minimum of ~ 34.46 at a depth of ~ 350m, and a

deep value of ~ 34.66. SPW is characterized by a strong maximum of ~ 35.50 at

a depth of ~ 100 — 150m, a weak minimum of ~ 34.61 at a depth of ~ 1500m,

and a deep value of ~ 34.65. Thus, it was shown that the deep temperature and

salinity distributions in the basin between the Sangihe Ridge and the topographic

rise at Dewakang Sill are shaped by the spilling of NPW, taking the western route,

over the Sangihe Ridge. The distributions in the basin between the Dewakang Sill

and the Flores-Banda ridge are formed by spilling of western route NPW over the

sill and flow of northern route NPW around the ridge. The deep temperature and

salinity distributions in the Morotai Basin and Maluku Sea are controlled by the

topographic rise just to the Pacific Ocean side of the Morotai Basin. The Banda

Sea deep temperature and salinity distributions are basically shaped by NPW taking

the northern route through the Lifamatola Passage and Seram Sea. It appears that

deep SPW is blocked by the northern and southern Halmahera Sills, so the deep

temperature and salinity distributions in the Seram Sea is basically formed by NPW

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 117

passing over the Lifamatola Sill. The relatively shallow Dewakang and Halmahera

Sills (~ 600m) essentially block the advection of temperature and salinity of those

layers in upstream regions of these sills that are below ~ 600m.

The analysis of observations presents some evidence of SPW in the upper 500m

in the Seram Sea, Banda Sea and to the north of the Lifamatola Passage (Aken et al.

1988; Gordon & Fine 1996). This seems to be the only contribution of SPW to the

formation of stratification of upper Banda Sea water. Unfortunately, due to some

reasons, discussed in Section 3, the model results of salinity distribution appeared

inconclusive on this point. Nevertheless, based on the direction of currents in the

upper layers of the Obi-Seram and Mangole-Buru sections and Lifamatola Passage

we can conclude that the model results do not contradict observations.

Note that we demonstrated a gradual evolution of NPW in the upper 500m along

the western route. No intrusion of water having substantially different temperature

and salinity characteristics could be found. Because the main part of the Indonesian

Throughflow passes through the Makassar Strait, we arrive at the conclusion that

basically it is NPW that shapes the temperature and salinity structure of the upper

part of the Indonesian Throughflow, as was stated by Wyrtki (1961); Gordon (1986);

and Gordon & Fine (1996).

There are no substantial structural changes of potential temperature and salinity

distributions between seasons, though values of some parameters of temperature and

salinity distributions (e.g., magnitudes of maxima and minima) can change. Thus,

our conclusions regarding the formation of deep temperatures and salinities in the

Seram and Banda Seas based on August data remain valid throughout the year.

It is important to stress that the main structure of the observed distributions of

temperature and salinity, namely the monotonic decrease of temperature with depth

and the presence of a maximum and minimum in salinity profiles, is reproduced very

well throughout the entire model domain. In the upper ocean (0-500m) the model

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 118

and observed profiles are very similar. In the deeper ocean temperature deviations

can reach 1 °C while salinity deviations are essentially on the order of several 0.01,

although in some layers, e.g., the layer of salinity minimum at the NGCC port, they

can reach as much as 0.1. The total transports of internal energy (heat) and salt

mass through the Lombok and Ombai Straits, and Timor Passage in August and

February are in reasonable agreement with published observed and simulated data.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 119

CHAPTER 4

THE ANALYSIS OF DISTRIBUTIONS OF TURBULENCE

CHARACTERISTICS BASED ON THE MODEL RESULTS

4.1. Introduction

Features of mixing are analyzed based on results of the integration of a regional

model of the Indonesian Seas using the Mellor-Yamada 2.5 turbulence model. Tur­

bulent mixing in the ITF system is well known to be important, see, for example,

the discussion of Gordon (2005). Two outstanding questions concerning the ITF and

the Indonesian Seas region are: i) how are T, S characteristics of thermocline NPW

and SPW transformed across the Indonesian seas and ii) how are deep overflows in

the Indonesian Seas sustained? The trivial answer to these questions is, in a word,

mixing. The objective of this chapter is to discuss systematically turbulence charac­

teristics within the ITF system. To the best of our knowledge, such a discussion for

the ITF region has not previously been given in the literature.

Turbulence measurements in the Indonesian Seas are a very scarce commodity.

The first discussion of turbulence in the Indonesian Seas we are aware of was by

Stommel & Fedorov (1967) who discovered microstructure in temperature and salin­

ity profiles from STD recorders near Timor and Mindanao. Alford et al. (1999)

measured microstructure in the Banda Sea near 100 m depth over a 2 week period.

Some investigators, such as Gordon et al. (2003), Van Aken et al. (1988), Berger

et al. (1988) and Bennekom (1988) have used simple models to estimate turbu­

lent diffusivities in the Indonesian Seas. Thus, the need to model turbulence in the

Indonesian Seas is urgent.

The results show that turbulent kinetic energy (TKE) is large at topographic

features such as sills. TKE is small in the interior ocean except in areas of strong

currents in the upper thermocline. Values are large near the surface dire mostly to

the monsoon winds but this effect is seen only to a depth of ~ 50m. Surface currents

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 120

are also quite energetic. Diffusivities are large in regions of high turbulence kinetic

energy. Model results seem realistic when compared with available observations and

other models. In the thermocline, large diffusivities may be sufficient to alter the

signature of NPW and SPW in the Indonesian Seas, while, at deep sills, they may be

sufficient to maintain the temperature gradient across sills thereby sustaining deep

overflows.

The material presented is as follows. The Mellor-Yamada 2.5 model is briefly

described in section 2, turbulence model results are presented in section 3 and a

discussion and conclusions are presented in section 4.

4.2. The Turbulence Parameterization Scheme:

The Mellor-Yamada (M-Y) 2.5 Model

The M-Y turbulence model has demonstrated skill in solving a wide range of

ocean and atmosphere turbulence problems. The full model consists of i) the energy

redistribution hypothesis of Rotta (1951), ii) the Kolmogorov (1941) hypothesis of

local, small scale isotropy and iii) adding closure expressions for third order moment

turbulent velocity diffusion and turbulent pressure diffusion terms. The Rotta and

Kolmogorov hypotheses lead to the definition of four length scales assumed to be

proportional to a master length scale. The resultant five constants are thus unam­

biguously related to measured laboratory turbulence for neutral flows. The M-Y 2.5

version is derived from a hierarchy of model versions by a process of formula sim­

plification based on assumed small departures from isotropy and application of the

boundary layer (BL) approximation. The M-Y 2.5 is a very popular second-moment

turbulence closure scheme widely used in research and operational models (see, e.g.,

Mellor & Yamada (1974, 1982), Mellor (2001), Galperin et al. (1988), Mellor (2004)).

The model solves for twice the turbulence kinetic energy (TKE), q2, and the master

turbulent length scale, /, as follows:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. dqHD 8Uq2lD dVq2lD duq2l d \Kq dq2l dt ^ dx dy ^ da da D da

2 Do3 ~ - —£-W$t(GH) + F,. (4.2) t> 11

The terms on the left side of (4.1) are the tendency and advective terms, and the

terms on the right hand side are the vertical turbulence diffusion, shear production,

buoyancy production, dissipation, and horizontal diffusion terms, respectively. D

is total depth of the water column; U and V are horizontal velocities; oj is the

POM vertical velocity; a is the vertical coordinate; Km is the calculated vertical

diffusivity for momentum (eddy viscosity); KH is the calculated vertical diffusivity

for temperature and salinity; Kq is the calculated vertical turbulence diffusivity and

is proportional to KH] g is gravity; p0 is the reference density; is the vertical

density gradient minus the adiabatic lapse rate; B\ is an empirical constant; and

I is the turbulence master length scale. In (4.2), W = 1 + F 2( ^ ) 2 where L^1 =

(■rj — z)~l + (H — z)~x is a wall proximity function and Ei,E2 and E?t are non-

dimensional constants.

1.0 if Gh > 0

% G h ) = 1.0 - 0.9M if G hc < G h < 0

0.1 if Gh < Ghc

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 122

where Gh = is a Richardson number Ghc — -6.0 is a constant and M —

(GH/Gh c Y ^2■ For weak turbulence, Gh < G hc , the dissipation is small. Note that

although equation (4.2) is the most empirical part of the model, it is used because it

has given reasonable results for multiple turbulent regions of the ocean. The solutions

to (4.1) and (4.2) above, lead to expressions of the turbulent diffusivities, KM and

I(a. where

Km — qISm, Kh = qlSn- (4.3)

The coefficients Sm and Sh are stability functions of Gh and are given by

6^! S h [1 — (3A2R2 + 18^4i^42) G h ] — -42 1 - 9(67 H, (4.4) b T

Sm [1 — §A\A2Gh\ — Sh [(1842 + G/r]

= Al [l-3Cl -6A lB1]^(GH). (4.5)

The five constants in (4.4) and (4.5) were evaluated from near-surface turbulence data

and decaying homogeneous turbulence data. They are (Mellor & Yamada (1982)):

(A\, Bi, A2, B2,C\) = (0.92,16.6,0.74,10.1,0.08). The stability functions Sm and

Sh limit to infinity as Gn approaches the value 0.0288, but this is not physically

possible since stratification would have been destroyed by the very large diffusivity

values associated with very large values of Sm and Sh - We note that a background

diffusivity of 2 .10- 5m2/,s is added to KM and Kh to represent internal waves and other mixing processes not modeled by the M-Y scheme. Solutions to (4.3), (4.4) and

(4.5) result in values of KM and KH that are used to solve the momentum and heat

and salinity equations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 123

4.3 Results The results are taken from the 1st of August during the South Eastern Monsoon

(SEM) after 10 years of model integration.

In deep water, large values of q2, KM and K h are seen near topographic features

such as sills. Values are greatest adjacent to the sills but patterns are complicated

and perhaps a little noisy. Maximum values of q2 are ~ 10-1 — lm 2/ s 2. The values

are greatest at the sills and along sloping topography but significant values can

extend in both upstream and downstrean directions. At deep sills large values of q2

extend above sill depth on the downstream side. In the interior ocean, away from

and above sill depth, q2 is reduced to less than 10_ 8 m2/s2. Patterns of KM and

Kh are very similar to q2. Maximum values of K m and KH are ~ 10-2 — lm 2/s

at the sills and along topography, and can be as large as 10_3m2/s for significant

distances downstream and upstream of the sills. In the thermocline q2 and the

diffusivities are generally small but increase noticeably where currents are known

to be strong with maximum values of q2 ~ 5.10~3m 2/s 2, K m ~ 5.10~2m2/s and

Kh ~ 5.10“3m 2/s. In the upper mixed layer (UML) large q2 and diffusivities are

generally confined to the upper 50m with maximum values of q2 ~ 10~3m 2/,s2 and

Km and K h ~ 5.10~2m 2/s. Turbulence master length scale, I, is large and generally

constant in the interior (~ 100m) and is small in the boundary layers. Stability

parameters, Sm and Sh , are greatest in the boundary layers where they approach 1.

They are generally small in the interior.

The results are given in Figures 4.1 to 4.6. Figure 4.1 shows the locations of

sections and profiles presented and discussed. The sections focus on 4 areas over

complex topography. Results across the North Sangihe Ridge area are presented in

Figure 4.2 and include (a) a section ( 1-2 in Figure 4.1) of q2] (b) KM', (c) KH; (d)

profiles of I at points across the section; and (e) and (f) profiles of stability factors

Sm and Sh , respectively. Note that the scale on the colorbar in the sections is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure 4.1. Locations of sections presented in the results. Section 1 - 2 is taken across the Northern Sangihe Ridge system from the Sulawesi Sea in the west to the Pacific Ocean in the east; section 3 - 4 is taken across the Dewakang Sill from the Southern Makassar Strait in the north to the Flores Sea in the south; section 5 -6 is taken across the Lifamatola Sill from the South Maluku Sea in the north to the North Banda Sea in the south; and section 7 - 8 is taken across the Halmahera Sea from the Obi-Seram passage in the south to the Pacific ocean in the north. The dots show the locations of profiles of 1 ,S m and SH.

logarithmic and profiles are shown on er-surfaces. All results are from August of

the seasonally varying experiment after 10 years of the model run. Figures 4.3 -

Figures 4.5 give the results across the Dewakang Sill, Limatola Sill and Halmahera

Sea, respectively. Figure 4.6 shows the effect of the wind.

4-3.1. North Sangihe Ridge Section

Across the North Sangihe Ridge (Figure 4.2) values of q2 in the vicinity of bottom

topography are ~ 10~3m 2/s2, or greater. The entire Karakelong Basin below the

depth of the Sangihe ridge, to the west, and the Karakelong Ridge, to the east, shows

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 125

q2 greater than 2.10“4m 2/ s 2. Below 1000m values generally greater than 2.10™4m 2/ s 2

extend 150 to 200 km downstream of the North Sangihe Ridge and into the Sulawesi

Sea. Upstream of the Karakelong Ridge, in the Pacific Ocean, q2 generally becomes

small within 100/cm of the sloping topography. In the thermocline, above ~ 1000m,

values of 10“ 8 m2/s 2 or less are seen except around the Karakelong Basin and ridge

where values are greater and maximum values of ~ 2.10_3m 2/ s 2 are found in the

upper 300m. In the UML values of ~ 2.10“ 4m 2/,s2 are confined to the upper 50m.

The spatial pattern in K m and Kh is very similar to q2. Near topography values

of Km are ~ 10- 2m2/s and often increase to 10” 1 to 1 rri2/s at the slope. Km values

are ~'5.10_3m2/s in the Karakelong Basin. Below 1000m, values of KM ~ 5.10“ 4

are seen downstream of the North Sangihe Sill. Km decreases to the background

value of 2.10“5m2/s away from topography. K m is small in the thermocline except

near the Karakelong Basin and Ridge where values are greater and reach 10- 2m2/s

between 200 and 300m. Values of KH are generally an order of magnitude less than

Km except in the UML (upper 50m) where both are ~ 5.10_ 3ra2/s.

The master length scale, I, is generally large in the interior ~ 1.00m and small

in the BLs, though the interior value is reduced at Karakelong Ridge (black profile).

Profiles of Sm and Sh have similar characteristics with maximum values approaching

1 in the UML and bottom boundary layer (BBL) and negligible values in the interior.

4-3.2. Dewakang Sill Section

Across the Dewakang Sill (Figure 4.3) q2 values of 2.10~ 4m 2/s 2 or greater are

seen. These values extend to 200m above the sill and increase to at least 10~ 3m 2/ s 2

along the passage between the sill and the Flores Sea. Along sloping topography,

values are generally reduced to background diffusivities on the downstream side to­

wards the Flores Sea, while patches of extremely high q2 ~ 10_2m 2/s 2 are seen on

the upstream side, South Makassar Strait. Large values are seen near the bottom

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 126

in the Flores Sea. Patches of high q2 extend through the water column at the De­

wakang Sill. Values in the UML are ~ 2.10_ 4m 2/s2. Away from topography values

are reduced to ~ 10“8 m2/s 2.

The pattern in KM and KH is very similar to q2. Maximum values of KM are

~ 5.10~3 to 10“2m2/s downstream of the sill to the Flores Sea. Values are generally

small along the upstream and downstream slopes and rapidly decrease to background

diffusivities. Values of Kh are generally an order of magnitude less than K m but are

similar in the UML where typical values are 10 ~3 or greater.

Profiles of Sm and Sh show large values in the boundary layers only, while I is

everywhere large outside the boundary layers.

4-3.3. Lifamatola Passage Section

The section across the Lifamatola Passage (Figure 4.4) extends from the South

Maluku Sea (Batjan Basin) across the Seram Sea into the Northern Banda Sea. In

deep water, downslope and downstream of the Lifamatola sill q2 values ~ 2.10~ 4m2/ s2

are seen. These values extend across the Seram Sea and into the North Banda Sea

where values can reach 10_ 2m 2/ s 2 in patches. Values of ~ 2.10_ 4m 2/ s 2 mostly ex­

tend throughout the lower 1000m of the water column between the sill and the North

Banda Sea. Upstream, in the Maluku Sea, very large values are seen in patches with

substantial areas having q2 ~ 2.10_4m 2/ s 2 or greater. Above ~ 1600m values are

small. In the UML q2 ~ 2.10_4m2/s 2.

The pattern in both K m and K h is similar to q2. Km values of 5.10~ 2m2/s

extend immediately downstream of the Lifamatola sill to the North Banda Sea.

Values greater than ~ 10“ 3m2/s are seen in the lower 1000m of the water column

in the Seram Sea, while values of ~ 5.10~ 4m2/s are seen to sill depth extending into

the North Banda Sea. Large values are also seen in the Maluku Sea, upstream of

the sill. Above ~ 1600m values reduce to the background value. Maximum Kh

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 127

values of 10 ~lm2/s or greater are seen downstream of the sill in the lower 500m of

the Seram Sea. In the UML values are as high as 5.10~ 2m 2/s. In the UML values

reach 5.10_ 2m2/s for both KH and KM.

Profiles of Sm and Sh show large values at the boundaries and small values in

the interior, while I is small in the boundaries and large outside.

4.3.4. Halmahera Sea Section

The Halmahera Sea section (Figure 4.5) extends from the Pacific Ocean to the

passage between Obi and Seram islands. Highest values of q2 ~ 5.10~3m 2/s 2 are seen

at the North Halmahera Sill in the upper 200?n and extend across the Halmahera

Sea to the southern sill where q2 is slightly less. Below sill depth, q2 ~ 2.10_4m 2/ s 2

throughout the Halmahera Basin. Downstream of the south sill along the slope in

the Seram Sea values are ~ 2.10“ 4m 2/s 2 and reach 5.10_ 3m 2/ s 2 in patches. Similar

patches are seen on the upstream slope in the Pacific Ocean. Away from topography

q2 is generally small. In the UML q2 ~ 2.10 4m2/s2.

Patterns in KM and KH are very similar to q2. KM is largest in the Halmahera

Basin where values ~ 5.10 ~~2 to 10- 1m2/s. In the upper 200m KM ~ 5.10- 2m2/s at

the Northern Halmahera Sill and ~ 10_ 3m2/s at the Southern Halmahera Sill. Km

is greater than 10~3m2/s in patches along the downstream slope. Kh is ~ 10-1 in

the Halmahera Basin, ~ 10 -3 in the upper 200m at the Northern Halmahera Sill and

~ 5.10-4 at the Southern Halmahera Sill. K m and Kh are ~ 5.10-2 in the UML.

Profiles of Sm and Sh show relatively large values outside the boundary layers,

particularly at the north sill (black profile) and I is also reduced in the interior,

particularly at the sills (black and blue).

4-3.5. Effect of the Wind on Turbulence Kinetic Energy (TKE)

Contour maps of q2 at different depths are given in Figure 4.6. The results show

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. distance(km)

5.10*2

10*2

1000 5.10*3

-S. 2000

3000 5.10*4 Karakelong Ridge 4000 North Sangihe Ridge Karakelong\ Basin 5000 5 .0 * 5 Sulawesi Pacific Ocean S ea 2 .0 * 5 200 250 300 350 400 distance(km)

Figure 4.2 Sections of (a) q2{m?/s2) (top) and (b) KM(m2/s) (bottom) across the North Sangihe Ridge. Locations of points 1 and 2 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 300m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 50 100 150 200 250 300 350 400 distance(km)

— i=138 j=221 i=146 j=221 10 1=154 j=221

$ 15

20

25

30. 100 200 300 400 500 600 700 800 900 1000 I ( m )

Figure 4.2 (cont.) Section of (c) KH(m2/ s) (top) and profiles of (d) l(m) (bottom) across the North Sangihe Ridge. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 300m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 130

10

1=138 j=221 [=146 j=221 j=15 4 j=221

20

25

30, 0.05 0.15 0.25 0.3 0.35 0.4

t i i ..... \ i ...... -r------r------1------1------

/ ...... _ ------..

-

“ ------i=138j=221 ~ i=146j=221 ------[= 154j=221

-

N ) I------1------_i—1------..... i 1______i i______i i______i i______j i ______i i______ti______0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5

Figure 4.2 (cont.) Profiles of (e) Sm (top) and (f) Sh (bottom) across the North Sangihe Ridge. Locations of profiles are given by the dots in Figure 4.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 100 200 300 distance(km)

IDe-2

S.IOe-3

IOe-3

£ 750 5.IOe-4

1000 Dewakang Sill 10e-4 1250 Flores S ea

1500 South M akassar Strait 2.108-5

150 200 250 300 350 distance(km)

Figure 4.3 Sections of (a) q2(m2/s 2) (top) and (b) KM{ni2/s) (bottom) across the Dewakang Sill. Locations of points 3 and 4 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 50 100 150 200 250 300 350 distance(km)

— — i=50 j=149 )=50 )=139 i=52 j=134

10

« 15

20

25

30. 10 20 3040 50 60 70 80 90 100 I (m )

Figure 4.3 (cont.) Section of (c) KH(m2/s) (top) and profiles of (d) l(m) (bottom) across the Dewakang Sill. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 133

10

= 5 0 j=149 S 15 =50 j=139 = 5 2 j=134

20

25

30, 0,05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 SM

1 1 1 1 ...... 1 1 1 1 1

— ------—

-

I ...... ------j=50 j=149 i=5Q j=139 ------i=52 j=134

-

v - ~ ------^

\ I______JI______...... - __ L1 ______...... 1.I______1I ______1I ______- /I______I t______It------...... i1 ------...... 1 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5

Figure 4.3 (cont.) Profiles of (e) Sm (top) and (f) Sh (bottom) across the De­ wakang Sill. Locations of profiles are given by the dots in Figure 4.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 50 100 150 200 250 300 350 distance(km)

5.t0e*2

2000

2500

3000 Lifamatola Strait 5.i0e*4

3500 Seram S e a 4000 Maluku S e a North Banda S ea 100 150 200 250 300 350 distance(km)

Figure 4.4 Sections of (a) q2(m2/s 2) (top) and (b) KM(,m2/s) (bottom) across the Lifamatola Sill. Locations of points 5 and 6 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 50 100 150 200 250 300 350 distance(km)

=138 j=156 10 =138 j=150 =131 j~144

20

25

30, 50 100 150 2 00 250 300 I (m)

Figure 4.4 (cont.) Section of (c) KH(m2/s) (top) and profiles of (d) l(m) (bottom) across the Lifamatola Sill. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the change in scale at 100m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 136

20

25

30, 0.05 0.15 0.25 0.3 0.35 0.4

10 =138 j=221 = 1 4 6 j=221 = 1 5 4 j=221 S> 15

20

25

30, 0.05 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5

SH

Figure 4.4 (cont.) Profiles of (e) Sm (top) and (f) Sh (bottom) across the Lifam­ atola Sill. Locations of profiles are given by the dots in Figure 4.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 137

|5.10e-3

2 .100-3

108-3 1000 5.108-4

1500 South Haimahera Sill North Haimahera Sill 2.108-4

2000 IOa-4

Haimahera Basin 2500 ,108-5

Seram S ea Pacific/ Ocean 2 .108-9 3000*- 100 150 200 250 350 distance(km)

IHf

_ 1000 5.108-3

South Haimahera Sill1500

North Haimahera Sill 2000 5.108-4

Haimahera Basin 2500

Seram S e a Pacific Ocean 2.10e-5 200 250 300 350 400 450 distance(km)

Figure 4.5 Sections of (a) q2(m2/s 2) (top) and (b) KM(m2/s) (bottom) across the Haimahera Sea. Locations of points 7 and 8 at the top of the figures are given in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the the change in scale at 150m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 50 100 150 200 250 300 350 400 450 distance(km)

10

$ 15

20

i=155 j=149 i= 162j= 156 25 —i=170j=164

30, 20 40 60 80 100 120 140 160 180 200 I (m )

Figure 4.5 (cont.) Section of (c) KH(m2/s) (top) and profiles of (d) l(m) (bottom) across the Haimahera Sea. Locations of profiles are given by the dots in Figure 4.1. Note the logarithmic scale on the colorbar. Note also the the change in scale at 150m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 139

i= 155j= 149 i= 162j= 156 £ 15 •j=170 j=164

20

30, 0.05 0.15 0.25 0.3 0.35 0.4

I 10 f =155 j=149 =162 j=156 £ 15 =170 j=164

20

25

30J- 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 S u

Figure 4.5 (cont.) Profiles of (e) SM (top) and (f) Sh (bottom) across the Haima­ hera Sea. Locations of profiles are given by the dots in Figure 4.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 140

i

Figure 4.6 Maps of (a) q2(m2/s 2) at the surface with climatological wind (top) and (b) no wind.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 141

Figure 4.6 (cont.) Maps of (c) q2(m2/s2) at 20m with climatological wind (top) and (d) no wind.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 142

Figure 4.6 (cont.) Maps of (e) q2(m2/s 2) at 50m with clirnatological wind (top) and (f) no wind.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 143

that the effect of the winds is felt to a depth of ~ 50m. Values of g 2 at the surface with

and without winds are given in Figures 4.6(a) and (b), respectively. These results

are given by the surfce boundary condition q2 = B \^ u 2 for experiment 7 (with wind)

where u2 is the friction velocity and q2 = 0 for experiment 8 (no wind). At 20m, (c)

and (d), values of ~ 10_3m 2/s 2 are seen around Haimahera, North Sangihe Ridge,

and Labani Channel for both experiments. Elsewhere, values of ~ 5.10~ 4m 2/s 2 are

generally seen for experiment 7 while values are very small for experiment 8 . At

50m, (e) and (f), the results are very similar. Values greater than 10~ 2ra2/s 2 are

seen near Haimahera, Sangihe and Makassar Strait. Values are small elsewhere. A

similar pattern is seen at 75 m and 100m (figures not shown).

4.4. Discussion and Conclusions

Turbulence and mixing are important in the Indonesian Seas. Of particular

interest are processes causing the transformation of thermocline water masses, NPW

and SPW, and processes maintaining deep overflows that are not well understood.

In addition, processes occurring at deep topographic sills in this region may help in

our understanding of the dynamics of deep overflows in general, which is important

for climate and ocean circulation studies.

In deep water in the Indonesian Seas, large values of q2 are found near sills and

around topography. Values are greatest along sloping topography but can be large for

substantial distances both downstream and upstream of the sill, though large values

generally extend for greater distances downstream. Values are generally small in the

thermocline except in areas of known currents. Patterns in K m and l\H are similar

to q2 and it appears that the turbulence mean velocity, q, is the major contributor

to large diffusivities. Large diffusivities in the upper thermocline in areas of known

currents are important for the diffusion of NPW and SPW, and to the maintenance

of overflows at sills and around topogaphy in the Indonesian Seas. The master length

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 144

scale, I, is small in the boundary layers and generally large in the interior, while the

stability parameters Sm and Sh are large in the boundary layers and generally small

in the interior. Sm and Sh can contribute to large diffusivities in the interior where

currents are strong, e.g., at the Haimahera Sills and Karakelong. However, I does

not contribute to large diffusivities since its value decreases where diffusivities are

large.

Distributions of KM and KH are very different at the four sills discussed, and

give an indication of the complex flow dynamics at the sills. The Sangihe Ridge

and the Lifamatola Sill are alike in that strong bottom overflows carry water into

the downstream basin, the Sulawesi and Seram Sea, respectively. Large values of

q2 associated with the overflows result in high diffusivities downstream of the sill.

These large diffusivities can mix warm water down to sill level thus ensuring the

temperature (pressure) gradient across the sill is maintained. At Dewakang, a process

similar to both Sangihe and Lifamatola occurs at the sill, ensuring the maintenance

of the overflow. However, large diffusivities are not seen along the sloping topography

of the Flores Sea and it is not clear how deep the overflow can penetrate or mix.

The Haimahera system is also very complex. Large values of q2, K M and KH are

seen at the Southern Haimahera Sill and along the bottom topography downstream.

A bottom current is also directed downstream across the southern Haimahera Sill

(Chapter 3).

Recall, from Chapter 3, that 0 and S distributions vary across the sill, i.e., 6 and

S are different on the upstream sloping topography, at the sill and on the downstream

sloping topography, respectively. Then fluid particles will change their 9 and S char­

acteristics as they upwell and downwell across the sill. The only possible mechanism

for this is diffusion. This seems quite plausible given the large diffusivities at the sill.

So, turbulent flows over steep topography are highly complex and variable at

and around the sills. Using available observations both Gordon et al. (2003) and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 145

Van Aken et al. (1988) calculated a mean vertical diffusivity for the Banda Sea

region below 300m of ~ 10“3m 2/s from Munk's (1966) abyssal recipe. Based on

radon profiles and deep silicate distribution, Berger et al. (1988) and Van Bennekom

(1988), respectively, found diffusivities as high as 2.10“~ 1m2/s in the deep Banda

Sea. Studies of the Faeroe Bank Channel Overflow show turbulent diffusivities with

values of 10~ 3m2/s to 10"2m2/s see, e.g., Saunders (1.990), Duncan et al. (2003),

Mauritzen et al. (2005). Mauritzen et al. (2005) also found that strong mixing

extends downstream at the Faeroe Bank Channel overflow while mixing is more

abrupt over the Denmark Strait overflow. Using the M-Y 2.5 turbulence model in an

idealized sloping basin, Mellor & Wang (1996) found diffusivities between 5.10“ 2m2/s

and 10_ 1m2/s in a layer extending 500m off the bottom along sloping topography.

These studies support our results of large q2 and diffusivities around topography.

The model diffusivities reduce to the background values of 2.10“ 5m2/s in the

interior ocean away from topography. Toole et al. (1994) using microstructure pro­

files found diffusivities of 5.10- 4m2/s near topography, decreasing to 10- 5m2/s away

from topography. Polzin et al. (1997) found values of l(V 5m2/s in the abyssal ocean

away from topography while Ledwell et al. (1993) also found values of 10~ 5m2/s in

the thermocline in the open ocean.

Large diffusivities are often seen in the upper thermocline in regions of strong

currents. In the Sangihe Ridge system, particularly large values are seen in the

eastern portion of the Karakelong basin and over the Karakelong Ridge to a depth

of 500m. These diffusivities are a result of the large q2 generated by the Mindanao

Current. In the Haimahera Sea system large diffusivities in the thermocline are due to the q2 generated by the New Guinea Coastal Current. Large diffusivities in the

thermocline are also seen at the Dewakang Sill and result from large q2 generated by

the strong southward current along the Makassar Strait.

Large diffusivities and q2 in the UML generally result from the SEM winds.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 146

However, energetic near surface currents entering the model domain through the

open ports, such as the Mindanao Current and the NGCC, and interior currents

such as in the Makassar Strait also lead to large values of q2.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 147

CHAPTER 5

SUMMARY

The circulation of the Indonesian Seas has been studied with a high resolution

regional model based on the POM. The model was developed as a research tool

to examine important physical processes occurring in the Indonesian Seas and the

different flow paths taken by the ITF. The primary objectives of this research were

to study: physical mechanisms controlling the splitting of the ITF between passages

and the effect of topography on flow paths taken; the effects of seasonal variation

of transports through the open ports; the influence of local winds on the regional

circulation; the role of topography in the formation of distributions of 0 and S in

the region with an emphasis on the interaction of NPW and SPW; and features of

mixing and turbulence having an important effect on 9, S and velocity distributions.

Physical mechanisms controlling the splitting of the ITF between passages and

the effect of topography on flow paths taken, the effects of seasonal variation of

transports through the open ports, and the influence of local winds on the regional

circulation were studied in Chapter 2 (objectives 1,2,3). All important topographic

features within the Indonesian Seas have been included in the developed regional

model which has horizontal resolution of ~ 10 km within a 250 x 250 grid and 29

cr-levels in the vertical. The model has four open ports to simulate inflow of North

Pacific Water from the Mindanao Current, inflow of South Pacific Water from the

New Guinea Coastal Current, outflow to the Pacific Ocean in the North Equator­

ial Counter Current, and outflow to the Indian Ocean in the ITF. Total transports

through the open ports and typical normal velocities at the ports are specified from

observations. Four experiments were discussed: seasonally varying and annual mean

port transports and normal velocities both with and without local winds. It was

shown that the portion of North Pacific Water entering the Indonesian Seas relative

to that leaving through the North Equatorial Counter Current port is fairly constant

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 148

throughout the year. Most of this water takes the western route through the Makas­

sar Strait. We have shown that the portion of South Pacific Water entering the

Haimahera Sea compared to that exiting in the North Equatorial Counter Current

varies considerably with the seasons.

It was shown that turning off the local winds does not substantially influence

transport through the main passages in the model domain. Surface circulation pat­

terns change substantially with the seasons. The comparison of the basic structure of

the model surface circulation and total transport values through the main passages

with observations appears satisfactory.

The role of topography in the formation of 6 and S distributions in the region

with an emphasis on the interaction of NPW and SPW, was examined in Chapter 3

(objective 4). This was considered by following the three major flow routes through

the Indonesian Seas: the western route from the Mindanao Port through the Sulawesi

Sea, along the Makassar Strait, across the Dewakang Sill, into the Flores Sea and

finally to the South Banda Sea; the northern route also from the Mindanao Port

across the Morotai Basin, the Maluku Sea, across Lifamatola into the Seram Sea,

the north Banda Sea and finally the South Banda Sea; and the eastern route from

the NGCC Port to the Northern Haimahera Sill, across the Haimahera Basin to the

Southern Haimahera Sill, and across the Obi-Seram passage to the Seram Sea. We

have shown that differences in stratification are caused by topographic features that

impede the advection of cold and salty water from an upstream basin to the basin

downstream. It was shown that in the upper 500m the ITF is primarily shaped by

NPW taking the western route through the Makassar Strait. Deep Banda Sea water

is formed by the overflow of NPW at the Lifamatola Sill, the deepest connection

between the Pacific and the Banda Sea. Finally, it was shown that the Haimahera

Sills block SPW from entering the Indonesian Seas below 500m. In the upper 500m,

SPW enters the Banda Sea through the Haimahera Sea and mixes with NPW in the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 149

Banda Sea region.

Features of mixing and turbulence having an important effect on 9, S and velocity

distributions, were outlined and analysed in Chapter 4 (objective 5). The turbulence

kinetic energy and the master turbulence length scale are calculated with the M-Y

2.5 turbulence parameterization model. It was shown that large turbulence kinetic

energies are found near topographic features such as sills. This energy results in

large diffusivities leading to mixing that can sustain overflows at sills. The results

show that the dynamics and processes are complicated and vary from sill to sill.

Known currents in the upper thermocline lead to strong diffusion of NPW and SPW

temperature and salinity signatures. The effect of the SEM winds on mixing is

limited to the upper 50m of the water column.

5.1. Future Work

One of the most interesting features of the Indonesian Seas circulation is the

deep overflow of water from a basin upstream to the neighboring basin downstream.

This is made possible by the complex bottom topography of the Indonesian Seas

region. The processes and flow dynamics of the overflow are very complicated and not

completely understood. We would like to investigate these overflows in more detail.

This could be done, perhaps, using very high resolution models around the area

of a known overflow and also with simple conceptual overflow models. This would

contribute to our knowledge of the overflow in general and also to our understanding

of the formation of deep water masses such as North Atlantic Deep Water, thereby

improving our understanding of the ocean circulation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 150

Appendix A

The smoothed bottom topography in the areas of the Sangihe Ridge, Labani

Channel, Dewakang Sill, Haimahera Sea, Lifamatola Passage, Lombok Strait, Ombai

Strait, Timor Passage, and the passage between the Irian Jaya and Seram Island is

shown in Figure Al. We now compare it with the unsmoothed detailed topography

given by Smith & Sandwell (1997), hereafter S & S:

a) The outstanding topographic features at the north and south of the Sangihe Ridge

are sills. At the northern sill, the location of the 1250m and 1500m isobaths agree

with S &S and the sill depth is 1300 — 1400m in both cases. At the southern sill, the

location of the 1000m and 1250m isobaths agree also with S &S and the sill depth

is between 1100 — 1200m in both cases.

b) The outstanding topographic feature at the Labani Channel is a sill. Locations

of the 1500m and 1750m isobaths agree with S&S. The sill depth is 1800 — 1900m

deep in our smoothed topography and 2000 — 2100m in S &S.

c) The outstanding topographic feature at Dewakang in the south Makassar Strait is

the Dewakang Sill. The locations of the 600m and 800m isobaths agree with S &S.

Our sill depth is 500 — 600m and S &S sill depth is ~ 600m. The passage to the west

of the Dewakang Sill and to the east of Kalimantan Island has depth 400 — 600m

and the location of the passage and depth range agree with S&S.

d) The outstanding topographic features in the Haimahera Sea are sills at the north­

ern and southern entrances to the sea itself. The 500m and 1000m isobaths agree

with S&S at both sills and the sill depths of 500 — 600m agree with S&S.

e) The outstanding topographic feature at the Lifamatola Passage is a sill. The

location of the smoothed 2000m and 1500m isobaths agree with S&S. Our smoothed

bathymetry gives a sill depth of ~ 1750 m whereas S&S give a sill depth of 1800 —

1900m.

f) The outstanding topographic feature between Burn and Seram Islands is a sill. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 151

location of 2000m and 1500m, isobaths agree with S&S. A sill depth of 1100 — 1200m

agrees with S&S. Isobaths between islands in the other 3 passages in this region agree

well with S&S.

g) The outstanding topographic feature between Irian Jaya and Seram Island is a

passage. The location of 1000m, 1500m and 2000m isobaths agree with S&S.

h) The outstanding topographic feature at the Lombok Strait is a sill. The location

of the 400m and 600m isobaths agree with S&S. The sill depth is ~ 400m in both

cases.

i) The outstanding topographic feature at Ombai Strait is a deep sill. The location of

1500m and 2000m isobaths in our smoothed topography agree with S&S. The depth

of our sill is 1600 — 1700m compared with 1400 — 1500m, in S&S.

j) The outstanding topographic feature between Timor Island and Australia is a deep

passage. The 1000m, and 1500m isobaths generally agree with S&S. The depth of

the passage is ~ 1500m in our smoothed topography whereas it is 1600 — 1800m in

S&S.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 135 140 145 30 40 50 60 5 10 15

155160165170175 (i)

130 140 150 160 180 110 115 i

Figure A.I. Some important features of the smoothed model bathymetry: (a) Sangihe Ridge with north and south sills; (b) South Makassar Strait with the De­ wakang Sill in the east and western passage near Kalimantan; (c) Lombok Strait;(d) Ombai Strait; (e) Timor Passage; (f) Haimahera Sea with north and south sills; (g) Seram Sea area with Lifamatola Passage to the northwest; Buru-Seram passage to the south; Obi-Halmahera and Obi-Seram passages to the east; and Mangole- Buru passage to the west; (h) Irian Jaya - Seram Passage; and (i) Labani Channel. Isobaths are in 1000‘s of meters.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 153

Appendix B

Horizontal boundary conditions at rigid walls for the depth-averaged motion :

1. Impermeability condition (normal component of velocity is equal to zero).

2. A special expression for the viscous flux of momentum in the direction normal to

the wall. This is sometimes called the ’half-slip’ condition. This condition has been

chosen because of the simplicity of its formulation in the C-grid.

Horizontal boundary conditions at rigid walls for the 3D motion

1. Impermeability condition (normal component of velocity is equal to zero).

2. A special expression for the viscous flux of momentum in the direction normal to

the wall (’half-slip’ condition).

3. Diffusive fluxes of temperature and salinity in the direction normal to the wall

are set to zero.

4. Diffusive fluxes of turbulence energy and length scale in the direction normal to

the wall are set to zero.

Vertical boundary conditions for the 3D motion

Surface:

1. The vertical turbulent flux of horizontal momentum is set equal to the prescribed

wind stress.

2. The kinematic boundary condition for the vertical velocity.

3. The prescribed temperature and salinity are set equal to the climatological value

at the surface

4. The turbulence energy is set proportional to the wind stress.

5. A special relation is assumed between the turbulence length scale and energy

at the second a-level. The turbulence length scale is set equal to zero at the first

cr-level.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 154

Bottom:

1. The bottom stress is set proportional to the square of the bottom velocity.

2. The kinematic boundary condition for the vertical velocity.

3. The vertical fluxes of heat and salt are set equal to zero.

4. The turbulence energy is set proportional to the bottom stress.

5. The turbulence length scale is set equal to zero.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 155

Appendix C

Table C.l calculated values e 5 rate of time variation (xlO8) -1.3x 10” 1 2.7x 10 ” 2 horizontal advective fluxes (xlO8) west Banda sea (i=90) including Timor passage, Ombai strait 9.7xl0 -1 2.5 and Flores sea, Q=7.3Sv north Banda sea (j=134), Q=-7.lSv -9.8X1CT 1 -2.4 east Banda section (i=145), Q=-0.2Sv 3.8x 10 ” 2 -7.4x10“ 2 Sum of horizontal advective fluxes 2.3xl(T 2 -9.4x 10” 3 horizontal diffusive fluxes (xlO8) west Banda sea (i=90) including Timor passage, Ombai strait -6.5xl0 ” 5 3.6x 10” 6 and Flores sea north Banda sea (j=134) -9.2xl0 ” 5 -3.4 xlO ” 6 east Banda section (i=145) 7.4xl(T 5 -5.4xl0 ” 6 Sum of horizontal diffusive fluxes -8 .2 x 10” 5 -5.2 xlO ” 6 vertical flux at the surface (xlO8) -6.3xl0 ” 2 -2.8X10 ” 1 nudging (xlO8) 1.8X10 ' 1 2.6 xlO ” 1 sum of all terms (xlO8) 2.6 x l 0” 3 7.8xl0 ” 3

Table C. The same as Table 3 but for different boxes in the region: C.l (above) the Banda Sea box, C.2 the Makassar Strait box and C.3 box to the north and east of Lombok Strait.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 156

Table C.2 calculated values e S rate of time variation (xlO8) -9.5xl0 “ 3 5.5xl0 -3 horizontal advective fluxes (xlO8) south of Makassar box (j=142), Q=9.5Sv 1.7 3.3 north of Makassar box (j=191), Q=-9.5Sv -1.7 -3.3 sum of horizontal advective fluxes 2.4x 10-2 4.8xlCT 3 horizontal diffusive fluxes (xlO8) south of Makassar box (j=142) 2.7xl0 ‘ 4 1.7xl0 “ 5 north of Makassar box (j=191) 2.6 x l 0“ 5 3 .8xl0 “ 6 sum of horizontal diffusive fluxes 2.9xl0 “ 4 2 .1 x l 0“ 5 vertical flux at the surface (xlO8) -4.6 xlCT 3 1.3xl0 “ 2 nudging (xlO8) 3.0xl0 -2 -2.4xl0 - 2 sum of all terms (xlO8) 4.0xl0 “ 2 -l.lx lO - 2

Table C.3 calculated values e S rate of time variation (xlO8) -3.6x10-2 l.lx lO ” 2 horizontal advective fluxes (xlO8) north of Lombok box (j=142), Q=-0.4Sv -3.5xl0 - 2 -1.2 x 10- ' east of Lombok box (i=41), Q=-1.7Sv -3.4 xlO -1 -6 .1 x 10- ' south of Lombok box (j=114), Q= 2.lSv 4.5x10-' 7.3x10“ ' sum of horizontal advective fluxes 7.7xl0 - 2 -2.5 x H r 3 horizontal diffusive fluxes (xlO8) north of Lombok box (j=142) -3.2x 10“ 5 -8.1 x 10"6 east of Lombok box (i=41) -3.1 xlO " 4 -5.9xl0 “ 5 south of Lombok box (j=114) -3.8 xlO " 4 1.2 x l 0“ 5 sum of horizontal of diffusive fluxes -7.1xl0 -4 -5.5x 10“ 5 vertical flux at the surface (xlO8) -2.5x10-2 -3.4xl0 - 2 nudging (xlO8) -3.2x10-2 2 .6 x 10-2

sum of all terms (xlO8) -1.7x10-2 l.OxlO -3

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 157

Appendix D

It is useful to write down the ’vertical’ velocity wpom used in the POM in the following form. First, consider cr-surfaces (surfaces at which a is constant). Their equation is

z = crD + r], D = H + r), (D.l) where H is the depth of the ocean and r) is the sea surface height. Introduce the unit normal vector to the o- surface,

1/2 dz dz ' dz ' ' dz ' , 1 , N + + 1 (D.2) N mxdx rriydy m Tdx mTdx

where mx and my are scale factors for the curvilinear orthogonal coordinate system x,y. Consider the velocity of an arbitrary point of the cr-surface and denote the normal component of this velocity by Va. Then it is easy to show that

wpom N (u • n„ - K,), (D.3) where u is the velocity of a fluid particle.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 158

REFERENCES

Aken, H.M. van, J. Punjanan, and S. Saimima. 1988. Physical aspects of the flushing

of the East Indonesian basins. Neth. J. Sea Res., Vol. 22, 315-339.

Aken, H.M. van, A.J. van Bennekorn, W.G. Mook, and H. Postma. 1991. Application

of Munk’s abyssal recipes to tracer distributions in the deep waters of the South

Banda basin. Oceanologica Acta, Vol. 14, 151-162.

Alford, M., M. Gregg, and M. Ilyas. 1999. Diapycnal mixing in the Banda Sea:

results of the first microstructure measurements in the Indonesian throughflow. Geo-

phys. Res. Letts., 26, 2741-2744.

Bennekom, A.J. van. 1988. Deep-water transit times in the eastern Indonesian

basins, calculated from dissolved silica in deep and interstitial waters. Proc. Snellius-

II Symp. Neth. J. Sea Res., Vol. 22, 341-354.

Bray, N.A., S. Hautala, J. Chong, and J. Pariwono. 1996. Large-scale sea level,

thermocline, and wind variations in the Indonesian throughflow region. J. Geophys.

Res., 101, 12239-12254.

Berger, G.W., A.J. Van Bennekom, and H.J. Kloosterhuis, 1988. Radon profiles in

the Indonesian archipelago. Neth. J. Sea Res., Vol. 22, 395-402.

Broecker, W.S. 1991. The great ocean conveyer. Oceanogr., 4, 79-89.

Burnett, W.H., V.M. Kamenkovich, A.L. Gordon, and G.L. Mellor. 2003. The

Pacific/Indian Ocean pressure difference and its influence on the Indonesian Seas

circulation: Part I - The study with specified total transports. J. Mar. Res., 61,

577-611.

Butt, J., and E. Lindstrom. 1994. Currents off the east coast of New Ireland, Papua New Guinea, and their relevance to regional undercurrents in the Western Equatorial

Pacific ocean, J. Geophys. Res., 99, 12503-12514.

Conkright, M.E., R. A. Locarnini, H.E. Garcia, T.D. O’Brien, T.P. Boyer, C. Stephens,

and J.I. Antonov. 2002. World Ocean Atlas 2001: Objective Analyses, Data Statis-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 159

tics, and Figures, CD-ROM Documentation. National Oceanographic Data Center,

Silver Spring, MD, 17 pp.

Cresswell, G., A. Frische, J. Peterson, and D. Quadfasel. 1993. Circulation in the

Timor Sea. J. Geophys. Res., 98, 14379-14389.

Cresswell, G.R., and J.L. Luick. 2001. Current measurements in the Halmahera Sea.

J. Geophys. Res., Vol. 106, 13945-13951.

da Silva, A. M., C. C. Young, and S. Levitus. 1994. Atlas of Surface Marine Data

1994, Volume 2: Anomalies of Directly Observed Quantities. NOAA Atlas NESDIS

7, U.S. Department of Commerce,NOAA, NESDIS.

de las Heras, M.M., and R. Schlitzer. 1999. On the importance of intermediate water

flows for the global ocean overturning. J. Geophys. Res., 104, 15515-15536.

Duncan, L.M., H.L. Bryden, and S.A. Cunningahm. 2003. Friction and mixing in

the Faeroe Bank Channel outflow. Oceanol. Acta, 26, 473-486.

Ffield, A., and A.L. Gordon. 1992. Vertical mixing in the Indonesian Thermocline.

J. Phys. Oc., 22, 184-195.

Ffield, A., and A.L. Gordon. 1996. Tidal Mixing Signatures in the Indonesian Seas.

J. Phys. Oc., 26, 1924-1937.

Ffield, A., K. Vranes, A.L. Gordon, and R. Dwi Susanto. 2000. Temperature vari­

ability within Makassar Strait. Geophys. Res. Letts., 27, 237-240.

Fieux, M., C. Andrie, E. Charriaud, A.G. Ilahude, N. Metzl, R. Molcard, and J.C.

Swallow. 1996. Hydrological and chlorofluoromethane measurements of the Indone­

sian throughflow entering the Indian Ocean. J. Geophys. Res., 101, 12433-12454.

Fieux, M., C. Andri, P. Delecluse, A.G. Ilahude, A. Kartavtseff, F. Mantisi, R.

Molcard, and J.C. Swallow. 1994. Measurements within the Pacific-Indian oceans

throughflow region. Deep-Sea Res. I, 41, 1091-1130.

Galperin, B., L.H. Kantha, S. Hassid, and A. Rosati. 1988. A quasi-equilibrium

turbulent energy model for geophysical flows. J. Atmos. Sci., Vol. 45, 55-62.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 160

Ganachaud, A. 1999. Large-scale oceanic circulation and fluxes of freshwater, heat,

nutrients, and oxygen. PhD Diss., MIT-WHOI Joint Prog., Woods Hole, MA,

pp.266.

Ganachaud, A., and C. Wunsch. 2000. Improved estimates of global ocean circula­

tion, heat transport and mixing from hydrographic data. Nature, Vol. 408, 453-457.

Ganachaud, A., and C. Wunsch. 2003. Large-Scale Ocean Heat and Freshwater

Transports during the World Ocean Circulation Experiment. J. Clim., Vol. 16,

696-705.

Godfrey, J.S. 1989. A Sverdrup model of the depth-integrated flow for the World

Ocean, allowing for island circulations. Geophys. Astrophys, Fluid Dyn., 45, 89-

112 .

Godfrey, J.S. 1996. The effect of the Indonesian Throughflow on ocean circulation

and heat exchange with the atmosphere: a review. J. Geophys. Res., 101, 12217-

12237.

Godfrey, J.S., and Y. Masumoto. 1999. Diagnosing the mean strength of the In­

donesian throughflow in an ocean general circulation model. J. Geophys. Res., 104,

7889-7895.

Godfrey, J.S., and J.L.Wilkin. 1995. Reply. J. Phys. Oceanogr., Vol. 25,1568-1570.

Godfrey, J.S., and T.J. Golding. 1981. The Sverdrup Relation in the Indian Ocean

and the effects of Pacific-Indian Ocean throughflow on Indian Ocean circulation and

on the East Australian Current. J. Phys. Oceanogr., Vol. 10, 11771-11779.

Godfrey, J.S., A.G.Hirst, and J.Wilkin. 1993. Why does the Indonesian Throughflow

appear to originate from the North Pacific? J. Phys. Oceanogr., Vol. 23, 1087-1098.

Goodman, P.J. 1998. The role of North Atlantic Deep Water Formation in an

OGCMs ventilation and thermohaline circulation. J. Phys. Oceanogr., Vol. 28,

1759-1785.

Gordon, A.L. 2005. Oceanography of the Indonesian Seas and Their Throughflow.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 161

Oceanography, Vol. 18, 4, 14-27.

Gordon, A.L. 2001. Interocean exchange. Chapter 4.7. In: Ocean Circulation and

Climate, G. Siedler, J. Church, J. Gould, Eds., Academic Press, 303-314.

Gordon, A.L. 1995. When is ”appearance” reality? A comment on why does

the Indonesian Throughflow appear to originate from the North Pacific. J. Phys.

Oceanogr., Vol. 25, 1560-1567.

Gordon, A.L. 1986. Interocean exchange of thermocline water. J. Geophys. Res.,

Vol. 91, 5037-5046.

Gordon, A.L., and R.D. Susanto. 2001. Banda Sea Surface Layer Divergence. Ocean

Dynamics, 52, 2-10.

Gordon, A. L., and R. D. Susanto. 1999. Makassar Strait transport: Initial estimate

based on Arlindo results. Mar. Tech. Soc. J., 32, 34-45.

Gordon, A.L., and R.A. Fine. 1996. Pathways of water between the Pacific and

Indian oceans in the Indonesian seas. Nature, Vol. 379, 146-149.

Gordon, A. L., and J. L. McClean. 1999. Thermohaline stratification of the Indone­

sian Seas: Model and observations. J. Phys. Oceanogr., Vol. 29, 198-216.

Gordon, A.L., A. Ffield, and A.G. Ilahude. 1994. Thermocline of the Flores and

Banda seas. J. Geophys. Res., Vol. 99, 18235-18242.

Gordon, A.L., C.F. Giulivi, and A.G. Ilahude, 2003(a). Deep topographic barriers

within the Indonesian seas. Deep-Sea Res. II, Vol. 50, 2205-2228.

Gordon, A.L., R.D. Susanto, and A. Ffield. 1999. Throughflow within Makassar

Strait. Geophys. Res. Letts., 26, 3325-3328.

Gordon, A.L., R. Dwi Susanto, and K. Vranes. 2003(b). Cool Indonesian throughflow

as a consequence of restricted surface layer flow. Nature, Vol. 425, 824-828.

Gouriou.Y., and J. Toole. 1993. Mean Circulation of the Upper Layers of the

Western Equatorial Pacific Ocean. J. Geophys. Res., 98, 22495-22520.

Haney, R.L.. 1971. Surface thermal boundary condition for ocean circulation models.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 163

Fluid Dynamics, Vol. 75, 91-106.

Kindle, J.C., G.W. Heburn, and R.C. Rhodes. 1987. An estimate of the Pacific

to Indian Ocean Throughflow from a global numerical model. Further Procgress in

Equatorial Oceanography., E.J. Katz and J.M. Witte, eds, Nova University Press,

pp. 317-321.

Kindle, J.C., H.E. Hurlburt, and E.J. Metzger. 1989. On the Seasonal and Inter­

annual Variability of the Pacific to Indian Ocean Throughflow. Proceedings of the

western Pacific international meeting and workshop on TOGA COARE, J. Picaut,

R. Lukas, T. Delcroix, eds, pp. 355-365.

Ledwell, J.R., A.J. Watson, and C.S. Law. 1993. Nature, 364, 701-703.

Lee, T., L Fukimori, D. Menemenlis, Z. Xing, and L.L. Fu. 2002. Effects of the

Indonesian Throughflow on the Pacific and Indian Oceans. J. Phys. Oceanogr., Vol.

32, 1404-1429.

Lindstrom, E., J. Butt, R. Lukas, and J.S. Godfrey. 1990. THE flow through

Vitiaz Strait and St. George’s Channel, Papua New Guinea. In: The Physical

Oceanography of Sea Straits, 171-189, NATO ASI Series, Ed.: L. Pratt, pp587.

Lindstrom, E., R. Lukas, R. Fine, E. Firing, S. Godfrey, G. Meyers and M. Tsuchiya.

1987. The Western Equatorial Pacific Ocean Circulation Study. Nature, 330, 533-

537.

Lukas, R., E. Firing, P. Hacker, P.L. Richardson, C.A. Collins, R. Fine and R. Gam­

mon. 1991. Observations of the Mindanao Current During the Western Equatorial

Pacific Ocean Circulation Study. J. Geophys. Res., 96, 7089-7104.

Lukas, R., T. Yainagata, and J.P. McCreary. 1996. Pacific low-latitude western

boundary currents and the Indonesian Throughflow.J. Geophys. Res., 101, 12209-

12216.

MacDonald, A.M., and C. Wunsch. 1996. An estimate of global ocean circulation

and heat fluxes. Nature, 382, 436-439.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 164

Marchesiello, P., J. M. McWilliams, and A. Shchepetkin. 2001. Open boundary

conditions for log-term integration of regional oceanic models. Ocean Modelling, 3,

1- 20.

Marotzke, J., R. Giering, Q.K. Zhang, D. Stammer, C.N. Hill, and T. Lee. 1999.

Construction of the adjoint MIT ocean general circulation model and application to

Atlantic heat transport sensitivity. J. Geophys. Res., 104, 29529-29548.

Marshall, J., A. Adcroft, C Hill, L. Perelman, and C. Heisey. 1997. A finite-volume

incompressible Navier-Stokes model for studies of the ocean on parallel computers.

J. Geophys. Res., 102, 5753-5766.

Mauritzen, C., J. Price, T. Sanford, and D. Torres. 2005. Circulation and mixing in

the Faeroese Channels. Deep-Sea Res. II, 52, 883-913.

Masumoto, Y., and T. Yamagata. 1996. Seasonal variations of the Indonesian

Throughflow in a general ocean circulation model. J. Geophys. Res., 101, 12287-

12293.

Mellor, G.L. and T. Yamada. 1974. A hierarchy of turbulent closure models for

planetary boundary layers. J. Atmos Sci., Vol. 13, No. 7, 1791-1806.

Mellor, G.L. and T. Yamada. 1982. Development of a turbulent closure model for

Geophysical Fluid Problems. Rev. Geophys. Space Phys., Vol. 20, 851-875.

Mellor, G. L., T. Ezer, and L.-Y. Oey. 1994. The pressure gradient conundrum of

sigma coordinate ocean models. J. of Atmospheric and Oceanic Technology, vol. 11,

1126-1134.

Mellor, G.L. 2001. One-dimensional ocean surface layer modeling, a problem and a

solution. J. Phys. Oceanogr., 31, 790-809.

Mellor, G.L. 2004. Users Giude for a three-dimensional, primitive equation, numer­

ical ocean model, Program in Atmospheric and Oceanic Sciences, Princeton Univer­

sity, pp.56.

Mellor, G.L., and X.H. Wang. 1996. Pressure Compensation and the Bottom Bound­

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 165

ary Layer. J. Phys. Oceanogr., Vol. 26, 2214-2222.

Miyama, T., T. Awaji, K. Akitomo, and N. Imasoto. 1995. Study of seasonal

transport variations in the Indonesian Seas. J. Geophys. Res., 100, 20517-20541.

Molcard, R., M. Fieux, and G. Ilahude. 1996. The Indo-Pacific throughflow in the

Timor Passage. J. Geophys. Res., 101, 12411-12420.

Molcard, R., M. Fieux, and F. Syamsudin. 2001. The throughflow within Ombai

Strait. Deep-Sea Res. I, 48, 1237-1253.

Molcard, M. Fieux, J.C. Swallow, G. Ilahude, and J. Banjarnahor. 1994. Low

frequency variability of the currents in the Indonesian Channels (Savu-Roti Ml and

Roti-Ashmore Reef M2). Deep-Sea Res., 41, 1643-1662.

Munk, W.H. 1966. Abyssal recipes. Deep-Sea Res., 13, 707-730.

Murray, S.P., and D. Arief. 1988. Throughflow into the Indian Ocean through the

Lombok Strait, January 1985-January 1986, Nature, 333, 444-447.

Murray, S.P., and D. Arief. 1990. Characteristics of circulation in an Indonesian

Archipelago Strait from hydrography, current measurements and modeling results.

In: The Physical Oceanography of Sea Straits, 3-23, NATO ASI Series, Ed.: L. Pratt,

PP587.

Murray, S.P., D. Arief, J.C. Kindle, and H.E. Hurlburt. 1990. Characteristics of

circulation in an Indonesian archipelago strait from hydrography, current measure­

ments and modeling results. The Physical oceanography of Sea Straits, L.J. Pratt

(ed.), 2-23.

Murray, S., E. Lindstrom, and J. Kindle. 1995. Transport Through the Vitiaz Strait.

WOCE Notes, 7, 21-23. Murtugudde, R., A.J. Busalacchi, and J. Beauchamp. 1998. Seasonal-to-interannual

effects of the Indonesian throughflow on the tropical Indo-Pacific Basin. J. Geophys.

Res., Vol. 103, 21425-21441.

Nielsen, M.H., L.J. Pratt, and K. R.Helfrich. 2004. Mixing and entrainment in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 166

hydraulically driven stratified sill flows. J. Fluid Mech., Vol. 515, 415-443.

Orlanski, I., 1976. A simple boundary condition for unbounded hyperbolic flows. J.

of Computational Physics. 21, 251-269.

Piola, A.R., and A.L. Gordon. 1984. Pacific and Indian Ocean Upper-Layer Salinity

Budget. J. Phys. Oceanogr., Vol. 14, 747-753.

Polzin, K.L., J.M. Toole, JR . Ledwell, and R.W. Schmitt. 1997. Spatial Variability

of Turbulent Mixing in the Abyssal Ocean. Science, 276, 93-96.

Potemra, J.T. 1999. Seasonal variations of upper ocean transport from the Pacific

to the Indian ocean via Indonesian Straits. J. Phys. Oceanogr., Vol. 29, 2930-2944.

Potemra, J.T., S.L. Hautala, and J. Sprintall. 2003. Vertical structure of Indonesian

throughflow in a large model. Deep-Sea Res. II, 50, 2143-2161.

Potemra, J.T., R. Lukas, and G.T. Mitchum. 1997. Large-scale estimation of trans­

port from the Pacific to the Indian Ocean. J. Geophys. Res., Vol. 102, 27795-27812.

Pratt, L.J., and P:A. Lundberg. 1991. Hydraulics of rotating strait and sill flow.

Annu.Rev. Fluid Mech., Vol. 23, 81-106.

Pratt, L.J., and M. Chechelnitsky. 1997. Principles for capturing the upstream

effects of deep sills in low resolution ocean models. Dyn. Atmos. Oceans, Vol. 26,

1-25.

Qiu, B.. and T.M. Joyce. 1992. Interannual Variability in the Mid- and Low-Latitude

Western North Pacific. J. Phys. Oceanogr., 22, 1062-1079.

Qiu, B., and R. Lukas. 1996. Seasonal and interannual variability of the North Equa­

torial Current, the Mindanao Current, and the Kuroshio along the Pacific western

boundary. J. Geophys. Res., 101, 12315-12330. Qiu, B., M. Mao, and Y. Kashino. 1999. Intraseasonal variability in the Indo-Pacific

throughflow and the regions surrounding the Indonesian Seas. J. Phys. Oceanogr.,

29, 1599-1618.

Qu, T., H. Mitsudera, and T. Yamagata. 1998. On the western boundary currents

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 167

in the . J. Geophys. Res., 103, 7537-7548.

Riel, P.M. van. 1943. The bottom water, introductory remarks and oxygen content.

-Snellius Exped. 1929-1930, Vol II, part 5, Chapter I. E.J. Brill, Leiden: 1-62.

Saunders, P.M. 1990. Cold outflow from the Faeroe Bank Channel, J. Phys. Oceanogr.,

20, 29-43.

Schiller, A., J.S. Godfrey, P.C. McIntosh, G. Meyers, and S.E. Wijffels. 1998. Sea­

sonal Near-Surface Dynamics and Thermodynamics of the Indian ocean and Indone­

sian Throughflow in a Global Ocean General Circulation Model. J. Phys. Oceanogr.,

Vol. 28, 2288-2312.

Schmitz, W.J. 1996a. On the World Ocean Circulation: Volume I. Some global fea­

tures / North Atlantic circulation. Woods Hole Oceanographic Institution, Technical

Report, WHOI-96-Q3, 141 pp.

Schmitz, W.J. 1996b. On the World Ocean Circulation: Volume 11. The Pacific and

Indian Oceans / A global update. Woods Hole Oceanographic Institution, Technical

Report, WHOI-96-08, 237 pp.

Schneider, N. 1998. The Indonesian Throughflow and the Global Climate System.

J. of Clim. , Vol. 11, 676-6889.

Schneider, N,, and T.P. Barnett. 4997. Indonesian throughflow in a coupled general

circulation model. J. Geophys. Res., Vol. 102, 12341-12358.

Shriver, J.F., and H.E. Hurlburt. 1997. The contribution of the global thermohaline

circulation to the Pacific to Indian Ocean Throughflow via . J. Geophys.

Res., Vol. 102, 5491-5511.

Smith, W. H. F., and D. T. Sandwell. 1997. Global sea floor topography from

satellite altimetry and ship depth soundings. Science, vol. 277, 1956-1962.

Sprintall, J., J.T. Potemra, S.L. Hautala, N.A. Bray, and W.W. Pandoe. 2003. Tem­

perature ansd salinity variability in the exit passages of the Indonesian Throughflow.

Deep-Sea Res. II, 50, 2183-2204.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 168

Sprintall, J., S. Wijffels, A. L. Gordon, A. Ffield, R. Molcard, R. Dwi Susanto, I.

Soesilo, J. Sopaheluwakan, Y. Surachman, and H. Van Aken. 2004. INSTANT: A

new international array to measure the Indonesian Throughflow. Eos 85(39):369.

Sprintall, J., A. Gordon, R. Molcard, G. Ilahude, N. Bray, T. Chereskin, G. Cresswell,

M. Feng, A. Ffield, M. Fiux, S. Hautala, J. Luick, G. Meyers, J. Potemra, D. Susanto,

and S. Wijffels. 2000. The Indonesian Throughflow: Past, Present and Future

Monitoring. Proceedings from the Sustained Observations for Climate of the Indian

Ocean Workshop/Perth, Australia, November 2000, IOC/CLIVAR Technical Report,

2000.

Stammer, D., G. Wunsch, R. Giering, C. Eckert, P. Heimbach, J. Marotzke, A.

Adcroft, C.N. Hill, and J. Marshall. 2003. Volume, heat and freshwater transports

of the global ocean circulation 1993-2000, estimated from a general circulation model

constrained by World Ocean Circulation Experiment (WOCE) data. J. Geophys.

Res., Vol. 108, 3007, 1-23.

Stommel, H., and K.N. Fedorov. 1967. Small scale structure in temperature and

salinity near Timor and Mindanao. Tellus, 19, 306-324.

Susanto, R.D., and A.L. Gordon. 2005. Velocity and transport of the Makassar

Strait throughflow. J. Geophys. Res., 110(001005), 1-10.

Toole, J.M., and B.A. Warren. 1993. A hydrographic section across the subtropical

South Indian Ocean. Deep-Sea Res. 40, 1973-2019.

Toole, J.M., E. Zou, and R.C. Millard. 1988. On the circulation of the upper waters

in the western equatorial Pacific Ocean. Deep-Sea Res., 35, 1451-1482.

Toole, J.M., K.L. Polzin, and R.W. Schmitt. 1994. Estimates of Diapycnal Mixing

in the Abyssal Ocean. Science, 264, 1120-1123.

Top, Z., A. Gordon, P. Jean-Baptiste, M. Fieux, A.G. Ilahude, and M. Muchtar.

1997. 3He in Indonesian Seas: Inferences on deep pathways. Geophys. Res. Letts.,

24, 547 - 550.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 169

Ueki, I., Y. Kashino, and Y. Kuroda. 2003. Observations of current variations off

the New Guinea coast including the 1997-1998 El Nino period and their relationship

with Sverdrup transport. J. Geophys. Res., 108, No. C7, 36-1 to 36-17.

Verschell, M.A., J.C. Kindle, and J.J. O’Brien. 1995. Effects of Indo-Pacific through­

flow on the upper tropical Pacific and Indian Oceans. J. Geophys. Res., Vol. 100,

18409-18420.

Vranes, K., A.L. Gordon, and A. Ffield. 2002. The heat transport of the Indonesian

Throughflow and implications for the Indian Ocean heat budget. Deep-Sea Res. II,

49, 1391-1410.

Wajsowicz, R. C., A. L. Gordon, A. Ffield, and R. D. Susanto. 2003. Estimating

transport in Makassar Strait. Deep-Sea Res. 11,50,2163-2181.

Warren, B.A. 1999. Approximating the energy transport across oceanic sections. J.

Geophys. Res., 104, 7915-7920.

Waworuntu, J.M., S.L. Garzoli, and D.B. Olson. 2001. Dynamics of the Makassar

Strait. J. Alar. Res., 59, 313-325. .

Wijffels, S., E. Firing, and J. Toole. 1995. The mean structure and variability of the

Mindanao Current at 8N. J. Geophys. Res., 100, 18421-18435.

Wyrtki, K. 1961. Physical oceanography of the Southeast Asian waters. Naga Report

2. Scripps Institution of Oceanography, 195pp.

Wyrtki, K. 1960. On the presentation of ocean surface currents. Int. Hydrogr. Rev.,

37, 111-128.

Yaremchuk, M., and T. Qu. 2004. Seasonal Variability of the Large-Scale Currents

near the Coast of the Philippines. J. Phys. Oceanogr., 34, 844-855.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 162

J. Phys. Oceanogr., 1, 241-248.

Hautala, S.L., J.Sprintall, J.Potemra, J.C. Chong, W. Pandoe, N.Bray, and A.G.

Ilahude. 2001. Velocity structure and transport of the Indonesian throughflow in

the major straits restricting flow into the Indian Ocean. J. Geophys. Res., 106,

19527-19546.

Hellerman, S., and M. Rosenstein. 1983. Normal monthly wind stress over the world

ocean with error estimates. J. Phys. Oceanogr., 13, 1093-1104.

Hirst, A.C., and J.S. Godfrey. 1993. The Role of Indonesian Throughflow in a Global

Ocean GCM. J. Phys. Oceanogr., Vol. 23, 1057-1086.

Ilahude, A.G., and A.L. Gordon. 1996. Thermocline stratification within the In­

donesian seas. J. Geophys. Res., Vol. 101, 12401-12409.

IF in, A. M., V. M. Kamenkovich, S. F. Kanaev, T. G. Zhugrina, and L. I. Lavr-

ishcheva. 1974. An experiment in constructing a smoothed bottom relief of the

World Ocean. Oceanology, 14, 617-622 (Translated from Russian).

Inoue, M., and S.E. Welsh. 1993. Modeling seasonal variability in the wind-driven

upper-layer circulation in the Indo-Pacific region. J. Phys. Oceanogr., 23, 1411-

1436.

Kamenkovich, V.M. 1977. Fundamentals of ocean dynamics. Elsevier Sci. Publ.

Co., 249pp. Translation from Russian: Osnovy dinamiki okeana. Gidrometeoizdat,

Leningrad, 1973, 240pp.

Kamenkovich, V.M., W.H. Burnett, A.L. Gordon, and G.L. Mellor. 2003. The

Pacific/Indian Ocean pressure difference and its influence on the Indonesian Seas

circulation: Part II - The study with specified sea-surface heights. J. Mar. Res., 61,

613-634.

Kashino, Y., A. Ishida, and Y. Kuroda. 2005. Variability of the Mindanao Current:

Mooring observation results. Geophys. Res. Letts., 32, L18611, 1-4.

Killworth, P.D., 1994: On reduced-gravity flow through sills. Gephys. Astrophys.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.