Download PDF
ads:
UNIVERSIDADE FEDERAL DE PERNAMBUCO-UFPE
CENTRO DE TECNOLOGIA E GEOCIÊNCIAS-CTG
DEPARTAMENTO DE OCEANOGRAFIA-DOCEAN
PROGRAMA DE PÓS-GRADUAÇÃO EM
OCEANOGRAFIA
Seasonal variability of the heat and mass transport
along the western boundary of tropical Atlantic
Marcus André Silva
Recife/Brasil
2009
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
Marcus André Silva
Seasonal variability of the heat and mass transport
along the western boundary of tropical Atlantic
Tese apresentada ao Programa de
Pós-graduação em Oceanografia da
Universidade Federal de Pernambuco como
requisito parcial para obtenção do título de
Doutor em Ciências, na área de
Oceanografia Física.
Orientador: Dr. Moacyr Araújo
Recife/Brasil
2009
ads:
S586s Silva, Marcus André
Seasonal variablity of the heat and mass transport along
the western boundary of tropical Atlantic / Marcus André
Silva. - Recife: O Autor, 2009.
xv, 121f.; il., gráfs., tabs.
Tese (Doutorado) Universidade Federal de Pernambuco.
CTG. Programa de Pós-Graduação em Oceanografia, 2009.
Inclui Referências.
1. Oceanografia. 2. Atlântico tropical. 3. Correntes de
borda oeste. 4. Bifurcação da CSE. I. Título.
551.46 CDD (22. ed.) UFPE/BCTG/2009-256
I
Para minha filha
Mariana.
II
Agradecimentos
Aos meus pais, à Renata e minha filhinha Mariana pelo companheirismo,
amor e motivação que me dedicaram neste período.
Ao amigo e professor Dr. Moacyr Araújo pela sua orientação, dedicação e
principalmente por acreditar sempre no meu trabalho.
Ao Dr. Jacques Servain pelo imenso suporte dado na interpretação dos
resultados obtidos neste meu trabalho. Sem o qual este trabalho não seria o
mesmo.
Ao Dr. Pierrick Penven pela presteza em colaborar com códigos
computacionais bem como, com contribuições importantes na compreensão
dos resultados numéricos obtidos.
A todos os professores do Departamento de Oceanografia da UFPE, que
sempre me motivaram a seguir na carreira de pesquisador.
À CAPES pela concessão de Bolsa de Pesquisa, sem a qual este trabalho não
teria sido possível.
Agradeço aos meus amigos do LOFEC, Dóris, Marcelo, Patrícia, Fabiana,
Rodolfo, Isaac e Márcio, pela amizade e companheirismo.
Agradeço a Márcio e Dóris (LOFEC) pelo apoio em diferentes etapas deste
trabalho e à Miriam Calland pelo empenho e dedicação na revisão gramatical
deste documento.
Agradeço a Christina, Tiago, Pedro, Marcelo, Adilma e todos os amigos dos
momentos de descontração, que tornaram esta jornada mais alegre.
A todos os funcionários do DOCEAN, pela presteza e amizade.
E a todos aqueles que, de uma forma ou outra, contribuíram para a
realização deste trabalho.
III
Resumo
O Atlântico tropical compreendido entre 20ºN e 20ºS apresenta-se hoje como
chave para o entendimento das mudanças climáticas na Terra. Duas regiões
despertam particular interesse: A banda equatorial do Atlântico onde o
sistema de correntes interage com forçantes térmicos superficiais e
ressurgência de Ekman, como a área sudoeste do Atlântico tropical (05°S-
25°S / 20°W-47°W), onde parte da Corrente Sul Equatorial (CSE) penetra
pela borda leste e contribui com muitas das correntes de fronteira oeste ao
longo da plataforma continental brasileira. Entretanto, a variabilidade da
dinâmica nestas regiões, que se mostra importante por sua contribuição
sobre o clima da região nordeste do Brasil, apresenta-se pouco estudada. O
presente trabalho investigou estes importantes sistemas do Atlântico tropical
utilizando o ROMS (Regional Ocean Model System). A primeira área de
estudo compreendida entre 20°S-20°N e 42°W–15°E, com resolução
horizontal de 1/6º e 30 camadas sigma (que acompanham o terreno).
Variações sazonais do transporte zonal, estrutura das correntes e
distribuição da TSM (Março e Agosto) obtidos numericamente foram
avaliados e comparados com dados: de literatura, experimentais do PIRATA e
observados por satélite. Os resultados desta simulação mostram que o
modelo é capaz de reproduzir os principais aspectos da Subcorrente
Equatorial (SE), Contra-corrente Norte Equatorial (CNE), Corrente do Golfo
(CG) e os ramos central e norte dos sistemas de Corrente Sul Equatorial
(cCSE/nCSE), em diferentes seções ao longo do equador. A comparação
entre a estrutura térmica nos primeiros 500 m simulada e do Programa
PIRATA mostra uma Camada de Mistura (MLD) bem reproduzida,
particularmente, a ressurgência que induz uma MLD mais rasa verificada
nas boias mais à leste durante o inverno austral até o final da primavera
austral. A evolução sazonal do sistema Piscina Quente do Atlântico Sul
(SAWP) – Língua Fria (Cold Tongue) foi bem representado, que é importante
para futuras previsões de variabilidade climática sobre as fronteiras
continentais da parte sudoeste do Atlântico tropical. Do lado sudoeste do
Atlântico tropical (05°S-25°S / 20°W-47°W), O ROMS (Regional Ocean Model
System) foi utilizado pela primeira vez nesta área para simular a circulação
oceânica utilizando uma malha de resolução horizontal de 1/12º com 40
camadas sigma, que acompanham o terreno, para resolução vertical. Para
avaliação preliminar da configuração do ROMS adotada foram analisadas as
distribuições superficiais e verticais de temperatura, além de calculadas as
evoluções sazonais da camada bem misturada e dos balanços, atmosféricos e
oceânicos, envolvendo a troca de calor dentro da camada bem misturada. A
ordem de grandeza das componentes oceânicas (principalmente a difusão
vertical e a advecção horizontal) é da mesma ordem de grandeza dos
forçantes atmosféricos e quase sempre opostos entre si, com alguma
diferença de fase e transporte dentro das camadas mais superficiais.
Resultados de variabilidade interanual foram comparados com os primeiros
dois anos de perfis de temperatura observados advindos dos três fundeios do
programa PIRATA-SWE (Projeto PIRATA, Extensão Sudoeste). A estrutura
térmica simulada nas camadas mais superficiais do oceano está em
concordância com os resultados obtidos in situ. Resultados de simulação
IV
apontam para uma larga e relativamente fraca CSE, composta por uma
sequência de núcleos não bem definidos e próximos a superfície. O
transporte que flui para oeste da CSE nos primeiros 400 m de profundidade
ao longo da seção que atravessa as boias PIRATA-SWE, calculado para
simulação do ROMS entre 2005-2007, apresenta um volume médio
transportado de 14,9 Sv, com um máximo observado em JFM (15,7 Sv) e um
mínimo durante MJJ (13.8 Sv). Os resultados de simulação indicam que em
2005-2007 o transporte para oeste da CSE foi modulado pela variabilidade
da componente zonal do vento. Três seções zonais, posicionadas do
continente até a posição da boia PIRATA, confirmam transporte mais intenso
da Sub-corrente do Norte do Brasil (SNB), fluindo para norte, e uma
diminuição no transporte da Corrente do Brasil (CB),que flui para sul,
durante maio de 2006 e maio de 2007, quando a bifurcação do ramo sul da
CSE alcança sua posição mais ao sul. Por outro lado, o máximo escoamento
da CB foi registrado durante janeiro de 2006, janeiro de 2007 e março de
2007, com um mínimo da SNB fluindo para norte em dezembro de 2005 e
outubro/dezembro 2006, correspondendo ao período em que a bifurcação do
ramo sul da CSE alcança sua posição mais ao norte (OND). A Elevação da
Superfície do Mar (ESM) e a Energia Cinética turbulenta (ECT) superficial
calculada a partir das simulações e dos produtos AVISO Rio05 apontam na
superfície para os mais altos níveis de energia de meso-escala ao longo do
ramo central da CSE e da SNB/CB. Resultados de modelagem ecológica
usando o modelo NPZD acoplado com o ROMS confirmam esta região como
uma área oligotrófica. Resultados do modelo ecológico são comparados com
SeaWifs dataset e a dinâmica e a produção primária são localmente
discutidos. Estes resultados preliminares disponibilizam mais informações
diante da complexidade da região de divergência da SCE e encoraja-nos a
conduzir estudos mais detalhados a respeito da dinâmica e do transporte de
massa nessa região utilizando o ROMS. Este trabalho também apresenta a
necessidade de continuação, ampliação e extensão vertical para o sistema de
observação PIRATA-SWE, especialmente com medidas de salinidade em mais
níveis de profundidade, além da instalação de medidores de correntes.
Palavras-chave: Atlântico tropical, modelagem matemática, conteúdo
térmico, bifurcação da CSE, correntes de borda oeste.
V
Abstract
The Tropical Atlantic comprised between 20ºN and 20ºS presents today an
important key to the knowledge of climatic changes in the world. Two areas
are of particular interest: the zonal band of the Equatorial Atlantic ocean
where the current system interacts with surface thermal forcing and Ekman
upwelling, and the Southwestern tropical Atlantic (05°S-25°S / 20°W-47°W),
where part of the South Equatorial Current (SEC) enters at its eastern
border, feeding many western boundary currents along the Brazilian
continental shelf. However, the long-term variability of the dynamics in these
regions, which is also important for their contribution to the climate over
northeastern Brazil, is largely unknown. This present work investigated
these important systems of the South tropical Atlantic using the Regional
Ocean Model System (ROMS). The first study area was comprised between
20°S-20°N and 42°W–15°E, with an horizontal grid resolution of 1/6º and 30
vertical sigma (terrain-following) layers. Seasonal variations of zonal
transport, current structure and SST distribution (March and August)
obtained numerically were evaluated and compared with: literature, PIRATA
array and satellite data. Results show that the model is capable of
reproducing the mean features of the EUC/NECC/GC and cSEC/nSEC
systems, at different sections along the equator. The comparison between the
simulated thermal structure, upper 500 m ocean layer, and PIRATA Program
dataset shows a very well reproduced Mixing Layer Depth (MLD). In
particular, the upwelling induced shallower MLD verified in eastern buoys
during austral winter to late spring. The seasonal evolution of the
Southwestern Atlantic Warm Pool (SAWP) – Cold Tongue equatorial system
was quite represented, which is important for future forecasting of climate
variability along the continental boundaries in Southwestern tropical
Atlantic. The second part of this work has the first use of ROMS to simulate
the ocean circulation in the Southwestern tropical Atlantic. This modelling
used a isotropic horizontal grid resolution of 1/12
°
and 40 sigma layers. For
preliminary evaluation of this ROMS configuration, surface and vertical
thermal structures were explored. Indeed, seasonal evolutions of the surface
mixed layer, atmospheric and oceanic balances involving the mixed layer
heat budget were calculated. The magnitude of oceanic components (mainly
the vertical diffusion and the horizontal advection) is in the same order as
the atmospheric forcing, and practically always opposes to it, with some local
and seasonal timing differences and mass transports within the upper levels.
Interannual variability results were compared with the first two-year series
of observed thermal profiles derived from the three PIRATA-SWE moorings.
The simulated thermal structure in the upper ocean layers agrees well with
in-situ data. ROMS simulations point out a broad and relatively weak SEC
flow composed of a sequence of more or less defined near-surface cores. The
westward SEC transport for the upper 400 m along the PIRATA-SWE section,
calculated from the ROMS simulation for 2005-2007, shows an average
volume transport of 14.9 Sv, with a maximum observed in JFM (15.7 Sv),
and a minimum during MJJ (13.8 Sv). ROMS results indicate that the 2005-
2007 seasonal near-surface westward SEC transport is modulated by the
zonal wind variability. Three zonal sections extending from the American
VI
continent to the PIRATA buoy sites confirm that stronger northward NBUC
transport and decreasing BC transport were achieved during May 2006 and
May 2007, i.e. at the time the sSEC bifurcation reached its southernmost
position. On the other hand, the maximum southward BC flow was verified
during January 2006, January 2007 and March 2007, with a minimum
northward NBUC flow in December 2005 and October/December 2006,
corresponding to the period when the sSEC bifurcation reached its lowest
latitude (OND). Sea Surface Height (SSH) and the surface Eddy Kinetic
Energy (EKE) derived from simulations and AVISO Rio05 product point out
the highest surface meso-scale activity (EKE 50 cm
2
s
-2
) along the cSEC
and NBUC/BC patches. Results of an ecological modelling using NPZD
model coupled with the ROMS confirm this region as an oligotrophic area.
NPZD outputs were compared with SeaWifs dataset and dynamic and
primary production are locally discussed. Preliminary results provide
additional ingredients in the complexity of the SEC divergence region, and
encourages us to conduct a more detailed exploration of the dynamics and
mass transport of this region using the ROMS. This also shows the need to
continue, extend, and vertically upgrade the observational PIRATA-SWE
array system, especially with more levels of salinity measurements and the
installation of current measurements.
Keywords: Tropical Atlantic, mathematical modeling, SEC bifurcation, heat
content, western boundary current.
VII
Figure List
Figure 1.1. Chart of study area with domain limited by 10°N-40°S and 20°E-
60°W with a schematic representation of mean zonal and meridional
currents: The upper ocean northward NBUC flow (red solid line), and the
deepest southward transport by the DWBC (blue dashed line). The sSEC
bifurcation at 100m depth (solid line); between 200m and 500m depth
(dashed line); between 500m and 1200m depth (pointed line), according with
Stramma and England (1999). ................................................................... 18
Figure 3.1. Region of model integrations, solid lines indicate domain of 1/6º
climatological equatorial modelling. Dashed lines indicate southwestern
tropical Atlantic domain limits of 1/12º horizontal grid. ............................. 31
Figure 4.1. Study area and model integration domain. The sections A, B and
C are dashed lines and the squares over the equator line are the PIRATA
mooring sites. Bathymetry is represented by gray shading contour. ........... 35
Figure 4.3. Zonal transport (u-component) obtained from model simulation
(ROMS) along Section A, 25°W. Dashed lines represent the isotherm 20°C.
Monthly-averaged values for March and August. ........................................ 38
Figure 4.4. Seasonal variability of EUC, NECC, SECc and SECn transports
at 25W° section, obtained in simulations. .................................................. 38
Figure 4.5. Zonal transport (u-component) obtained from model simulation
(ROMS) along Section B, 15°W. Dashed lines represent the isotherm 20°C.
Monthly-averaged values for March and August. ........................................ 40
Figure 4.6. Seasonal variability of EUC, NECC, SECc and SECn transports
at 15W° section, obtained in simulations. .................................................. 41
Figure 4.7. Zonal transport (u-component) obtained from model simulation
(ROMS) along Section C, 5°W. Dashed lines represent the isotherm 20°C.
Monthly-averaged values for March and August. ........................................ 42
Figure 4.8. Seasonal variability of EUC, NECC, SECc and SECn transports
at 5W° section, obtained in simulations. .................................................... 43
Figure 4.9. Comparison of the vertical temperature distribution provided by
ROMS (black dashed contours) with the three PIRATA buoys (colour shading)
extending from surface down to 500 m. Monthly-averaged distributions for
the PIRATA period 1997-2007. ................................................................... 44
Figure 4.10. Monthly-averaged SST obtained from ROMS model (above) and
from satellite observations (AVHRR-NOAA) (below) for March and August. . 45
Figure 5.1. Model domain (dashed lines) with the PIRATA-SWE locations
(filled triangles) and three sites located along the western boundary region
closer to the Brazilian edge (filled circles). Near shore 100m and 1000m
isobaths are plotted. .................................................................................. 48
Figure 5.2. Horizontal distribution of SST obtained from ROMS in mid-
September (a) and mid March (b). PIRATA-SWE locations (filled triangles)
and three sites located along the western boundary region closer to the
Brazilian edge (filled circles). ...................................................................... 50
Figure 5.3. Horizontal distribution of SST obtained from observed satellite
data on 15 September 2005 (a) and 15 March 2006 (b). PIRATA-SWE
locations (filled triangles) and three sites located along the western boundary
region closer to the Brazilian edge (filled circles). ........................................ 51
Figure 5.4. Comparison of the monthly mean vertical temperature
VIII
distribution (ºC) provided by ROMS (black dashed contours) with monthly
mean for the period of September 2005 to February 2008 of the three
PIRATA-SWE buoys (colour shading and white lines) extending from surface
down to 500m. The four vertical black thin lines correspond to September
15
th
, December 15
th
, March 15
th
and June 15
th
,. (a) 8°S-30°W; (b) 14°S-
32°W; (c) 19°S-34°W. ................................................................................. 53
Figure 5.5. September (a), December (b), March (c) and June (d) monthly
averaged vertical profiles of temperature 0-500m at the three PIRATA-SWE
sites provided by climatic ROMS and by PIRATA monthly averaged profiles
from September 2005 to February 2008 ..................................................... 55
Figure 5.6. Comparison of temporal evolution of seasonal heat content in
the mixed layer at 8°S-30°W (a), 14°S-32°W (b) and 19°S-34°W (c), provided
by PIRATA in-situ observations (dashed line) and ROMS (solid line) for the
period of September to August. .................................................................. 57
Figure 5.7. Seasonal evolution of SST (°C) (black dashed line), atmospheric
(red line) and oceanic (blue line) contributions to the local change of SST
(°C/Month) (black continuous line) provided by ROMS at 8°S-34°W (a), 8°S-
30°W (b), 14°S-38°W (c), 14°S-32°W (d), 19°S-38°W (e) and 19°S-34°W (f)
locations. ................................................................................................... 61
Figure 5.8. Mean annual volume transport averaged across three zonal
sections at 8°S (a), 14°S (b) and 19°S (c), and across the section along the
PIRATA-SWE array (d). Positive (negative) values indicated by solid (dashed)
white lines correspond to northward (southward) currents (panels 8a, 8b,
and 8c), while positive (negative) values indicated by solid (dashed) white
lines for the section along the PIRATA array correspond to eastward
(westward) currents (panel 8d). The three horizontal solid black lines indicate
the 24.5, 26.8 and 32.15 sigma-t values (in kg.m-3). ................................. 63
Figure 6.1. Comparison between ROMS-derived SST (a and c) and GHRSST-
PP-derived SST (b and d) for September 15
th
2005 and March 15
th
2006. ... 67
Figure 6.2. Comparison between simulated and observed temperature (
o
C)
variations for the upper 500m from September 2005 to June 2007 for the
three PIRATA-SWE locations at (a) 08°S-30°W, (b) 14°S-32°W and (c) 19°S-
34°W. Model-derived temperature is represented by black contours and the
PIRATA-SWE observed temperature appears as white contours and shaded
colors. ....................................................................................................... 71
Figure 6.3. Comparison of monthly averaged vertical profiles of temperature
for the first 500m between the PIRATA-SWE in-situ observations (dashed
line) and the ROMS simulation (solid line) for: (a) September 2005, (b)
December 2005, (c) March 2006 and (d) June 2006. Colours are associated
to each PIRATA-SWE buoy: 08°S-30°W (blue), 14°S-32°W (black) and 19°S-
34°W (red). ................................................................................................ 73
Figure 6.4. Comparison of temporal evolution of seasonal mixed layer depth
(MLD) at: (a) 8°S-30°W, (b) 14°S-32°W and (c) 19°S-34°W, provided by
PIRATA-SWE in-situ observations (dashed line) with ROMS (solid line) for the
period of September 2005 to June 2007. .................................................... 75
Figure 6.5. Averaged (2005-2007) SSH and EKE comparisons between
ROMS simulations and AVISO Rio05 data: (a) ROMS - SSH (cm), (b) ROMS -
EKE (cm
2
s
-2
), (c) RIO05+AVISO - SSH (cm) and (d) AVISO - EKE (cm
2
s
-2
). . 77
Figure 6.6. Transport function (Sv) for the modelled seasonal averaged
IX
(three months, 2005-2007) currents integrated from 600 m to the sea
surface for a) MJJ and b) OND. The 1000 m isobath is represented by a red
line. ........................................................................................................... 78
Figure 6.7. Model simulation of meridional current and transport values
averaged from September 2005 to July 2007 along zonal sections (0-1500 m)
at: (a) 8°S, (b) 14°S and (c) 19°S. Positive (negative) values indicated by solid
(dashed) lines correspond to northward (southward) currents. ................... 80
Figure. 6.8. Annual averaged (2005-2007) meridional current (cm s-1) and
transport (in Sv) obtained by ROMS for: (a) the along zonal section (0-1500
m) at 7°S. Positive (negative) values indicated by solid (dashed) lines
correspond to northward (southward) currents, (b) the depth averaged range
200-1000 m, where CFLOW is stronger ...................................................... 83
Figure. 6.9. Interannual variability of the North Brazil Undercurrent (NBUC)
transport along 8oS (from the surface to 400 m), and Brazil Current (BC)
transport along 19oS (from the surface to 400 m), obtained from ROMS
simulations for the period of September 2005 to July 2007. Positive
(negative) values indicated by the black (gray) line correspond to northward
(southward) transports. ............................................................................. 85
Figure. 6.10. Seasonal averaged (three months, 2005-2007) meridional
velocity (m s-1) obtained from the ROMS simulation for a) MJJ and b) OND.
The velocities are averaged within a 1° longitude band off of the Brazilian
coast. The white line is the contour of zero velocity that represents the
bifurcation of the sSEC. The white areas represent the Vitoria-Trindade
Ridge and Abrolhos Bank. ......................................................................... 87
Figure 7.1. Model domain (dashed gray lines) with the PIRATA-SWE
locations (filled triangles). Section along three PIRATA-SWE sites (solid lines).
Thin dashed lines are 100m and 1000m isobaths, respectively. ................. 94
Figure 7.2. Chlorophyll a distribution on surface to SeaWifs on: a) November
b) July and ROMS results on: c) November d) July ..................................... 95
Figure 7.3. Vertical salinity distribution of ROMS along PIRATA-SWE sites
on: a) November b) July. ............................................................................ 97
Figure 7.4. Vertical nitrate distribution of ROMS along PIRATA-SWE sites
on: a) November b) July. ............................................................................ 99
X
List of Abbreviations
AABW Atlantic Antarctic Bottom Water
AAIW Atlantic Antarctic Intermediate Water
AC Agulhas Current
ACC Antarctic Circumpolar Current
ACM Acoustic Current Meter
ADCP Acoustic Doppler Current Profile
AMSR-E Advanced Microwave Scanning Radiometer
BC Brazil Current
Brazilian CHM Brazilian Navy Hydrographic Center
CC Coefficient of Correlation
CLIVAR Climate Variability and Predictability Program
COADS Comprehensive Ocean-Atmosphere Data Set
cSEC central band of the South Equatorial Current
CTW Coastal Trapped Wave
CW Central Water
CW Coastal Water
DMSP Defense Meteorological Satellites Program
DWBC Deep Western Boundary Current
EOF Empirical Orthogonal Functions
ERA-40 Reanalysis of the Global Atmosphere for 45 years
ERS European Remote Sensing Satellite
ETOPO-2 Digital Terrain Data with 2’’ resolution
EUC Equatorial Undercurrent
Guinea C. Guinea Current
GC Gulf Current
GEOSAT Geodetic Satellite
GFDL Geophysical Fluid Dynamics Laboratory
GLOSS/IOC Global Sea Level Observing System
GOES Geostationary Operational Environmental Satellites
HDF-EOS Hierarchical Data Format - Earth Observing System
XI
IBGE Institute of Geography and Statistics
IO/USP Institute of Oceanography of São Paulo University
IRD Institut de recherche pour le développement
ISV Intraseasonal Variability
ITCZ Intertropical Convergence Zone
IWBC Intermediate Western Boundary Current
MLD Mixed Layer Depth
MOC Meridional Overturning Cell
MODIS Moderate-resolution Imaging Spectroradiometer
MOM Modular Ocean Model
MSLA Mean Sea Level Anomaly
MSW Maximum Salinity Water
NADW North Atlantic Deep Water
NBC North Brazil Current
NBUC North Brazil Undercurrent
NCAR National Center for Atmospheric Research
NCEP National Centers for Environmental Prediction
NEC North Equatorial Current
nSEC northern band of the South Equatorial Current
NOAA National Oceanic and Atmospheric Administration
OCCAM Ocean Circulation and Climate Advanced Modeling
Project
PCA Principal Component Analysis
POD Proper Orthogonal Decomposition
POP Parallel Ocean Program
QuikSCAT Quick Scatterometer
RCM Recording Current Meter
RMPG Geodetic Permanent Tide Gauge Network
ROMS Regional Ocean Modeling System
SAC South Atlantic Current
SACW South Atlantic Central Water
SACZ South Atlantic Convergence Zone
SASH South Atlantic Subtropical High
XII
SBB Southeast Brazil Bight
SCF Squared Covariance Fraction
SEC South Equatorial Current
SECC South Equatorial Countercurrent
SEUC South Equatorial Undercurrent
SICC Southern Intermediate Countercurrent
SMW Salinity Maximum Water
SODA Simple Ocean Data Assimilation
sSEC southern band of the South Equatorial Current
SSM/I Special Sensor Microwave
SST Sea Surface Temperature
STC Subtropical Cell
Sv Sverdrup
SVD Singular Value Decomposition
TMI Microwave Imager
TOGA-TAO Tropical Ocean Global Atmosphere - Tropical
Atmosphere Ocean
TRMM Tropical Rainfall Measuring Mission
TSW Tropical Surface Water
TW Tropical Water
uCDW upper Circumpolar Deep Water
WOA World Ocean Atlas
ZCAS Convergence Zone of the South Atlantic
XIII
Contents
AGRADECIMENTOS ................................................................................................. III
RESUMO .................................................................................................................. IV
ABSTRACT .............................................................................................................. VI
FIGURE LIST ........................................................................................................ VIII
LIST OF ABBREVIATIONS .................................................................................. XI
CONTENTS ........................................................................................................... XIV
CHAPTER 1 ............................................................................................................ 16
1. MOTIVATION AND OBJECTIVE ................................................................ 16
CHAPTER 2 ............................................................................................................ 23
2. OBSERVATIONAL AND DERIVED DATA ................................................. 23
CHAPTER 3 ............................................................................................................ 29
3. THE MODELLING TOOLS ........................................................................... 29
CHAPTER 4 ............................................................................................................ 34
4. THE PATHWAYS AND THERMAL STRUCTURE OF SURFACE
CURRENTS IN THE EQUATORIAL ATLANTIC OCEAN .................................. 34
CHAPTER 5 ............................................................................................................ 46
5. NEAR SURFACE TRANSPORT AND HEAT BUDGET IN THE
SOUTHWESTERN TROPICAL ATLANTIC USING A HIGH-RESOLUTION
NUMERICAL MODELLING SYSTEM .................................................................. 46
CHAPTER 6 ............................................................................................................ 65
6.1. THE HIGH-RESOLVED INTERANNUAL ROMS APPROACH ................. 66
XIV
XV
CHAPTER 7 ............................................................................................................ 92
7. REGIONAL BIOGEOCHEMICAL MODELLING IN THE
SOUTHWESTERN TROPICAL ATLANTIC ......................................................... 92
CHAPTER 8 .......................................................................................................... 100
8. CONCLUSIONS AND PERSPECTIVES .................................................... 100
REFERENCES ...................................................................................................... 108
Chapter 1 – Motivation and objective
Chapter 1
1. Motivation and objective
The equatorial Atlantic Ocean is a very complex area where zonal currents
and counter-currents coexist with strong ocean-atmosphere heat exchanges
as well as local upwelling of cold waters. It is not surprising that several
research programs have been proposed to study the transport and heat
variability in this area, as for example the Global Atmospheric Research
Program (GARP), the Atlantic Tropical Experiment (GATE), the Francais
Ocean Climat Atlantique Equatorial (FOCAL), and the Pilot Research Moored
Array in the Tropical Atlantic (PIRATA). Most of these efforts focus on air-sea
interaction processes related to climate change investigations.
The importance of sea surface temperature (SST) variability in the
tropical Atlantic, for example, as a response to the seasonal varying winds
has been the subject of numerous studies. SST in the tropics is largely
determined by upper ocean dynamics including advection. Therefore, it is
necessary to understand and model its influence in climate, as well as, to
know the surface currents and their variability (Lumpkin and Garzoli, 2005).
Different work-groups have presented results from modeling and
16
Chapter 1 – Motivation and objective
observations confirming that the tropical Atlantic is a critical region for
processes associated to the Meridional Overturning Circulation (MOC), such
as cross-equatorial exchanges, and sea surface temperature variability that
impacts on climate variability of the coupled tropical ocean/atmosphere
system - see Stramma and Schott (1999), Lee and Csanady (1999a,b),
Stramma et al. (2003), for an overview. In the tropical ocean, cold
thermocline waters of non-equatorial origin from both hemispheres are
transformed into warm water masses via entrainment and subsequent
surface heat flux, and then escape from this region northward as well as
southward, although the net heat transport is northward. The tropical
Atlantic warm water pool stores heat during May-October and permits heat
escape in November-April.
The warm component of the ‘‘conveyor belt’’ in the equatorial Atlantic
is sensitive to this global circulation cell, but also to smaller horizontal scale
movements related to the subtropical subduction and to the equatorial
recirculation cell, for which Ekman divergence partially compensates
geostrophic convergence (Blanke et al, 1999). According to Stramma and
Schott (1999), the water mass exchange, including net northward meridional
heat transport across the equator, occurs on three layers (surface, central
and bottom). In the surface layer, it is accomplished by warm Tropical
Surface Water (TSW). In the central layer, Antarctic Intermediate Water
(AAIW) and upper Circumpolar Deep Water (uCDW) move northward in the
upper 1200 m and is compensated by cold North Atlantic Deep Water
(NADW), which moves southward between 1200 and 4000 m. In the bottom
layer, the northward transport of Antarctic Bottom Water (ABW) also carries
a small amount of cold water into the northern hemisphere.
The main zonal currents that act on the equatorial surface layer are
the North Equatorial Current (NEC), the North Equatorial Countercurrent
(NECC), the South Equatorial Current (SEC), and the Equatorial
Undercurrent (EUC) near the equator. These standards were more recently
detailed from drifter observations as analyzed by Lumpkin and Garzoli
(2005). In the South tropical Atlantic, the SEC flows westward on the upper
ocean circulation. Molinari (1982) described the SEC as divided into three
17
Chapter 1 – Motivation and objective
bands in the tropical Atlantic. The northern branch of SEC was called
northern SEC (nSEC), the flow between the SEUC and the SECC is known as
central SEC (cSEC), and finally, flowing south of the SECC, the southern
SEC (sSEC). On the western boundary of the tropical Atlantic, the sSEC
bifurcates into the North Brazilian Undercurrent (NBUC) that flows
northward, and the Brazilian Current (BC) that flows southward along the
Brazilian coast. In the beginning of the northern Brazilian shelf, near 4ºS,
the NBUC joins the cSEC to form the shallower NBC. After crossing the
equator, a part of NBC retroflects and contributes to eastward flows of the
NECC and EUC. Therefore the west border of the tropical Atlantic, near 35°
W, has particular importance for the mixture of water masses because it
contains a large portion of the Atlantic Meridional Overturning Circulation
(MOC) passing through it (Schott et al., 2003).
( Veleda, 2008)
Figure 1.1. Chart of study area with domain limited by 10
°
N-40
°
S and 20
°
E-60
°
W
with a schematic representation of mean zonal and meridional currents: The upper
ocean northward NBUC flow (red solid line), and the deepest southward transport by
the DWBC (blue dashed line). The sSEC bifurcation at 100m depth (solid line);
between 200m and 500m depth (dashed line); between 500m and 1200m depth
(pointed line), according with Stramma and England (1999).
Inside the southwestern border of the tropical Atlantic, the Northern
Atlantic Deep Water (NADW), originated by the cold water plunged from the
North Atlantic, which runs towards the South Atlantic, are relatively well-
18
Chapter 1 – Motivation and objective
known (Arhan et al., 1998; Stramma et al., 2005). However, in the southern
hemisphere, the formation of shallow warm water by subduction, and its
northward transfer, to close the heat budget of MOC across the equator, is
less understood. The southern Atlantic Ocean presents prime importance for
global climate change due to the number of key zones where oceanic signals,
on various scales (from intra-seasonal to decadal), crosses through. In the
southern subtropical Atlantic basin, mainly subjected to both cyclonic and
anticyclonic gyres strongly controlled by the surface wind (Stramma and
Schott, 1999; Lumpkin and Garzoli, 2005), the following key zones can be
listed, from South Africa to South America successively:
(i) The Agulhas Current off South Africa, coming up from the
Indian Ocean and propagating northwards while skirting the
African coast;
(ii) The Benguela Current system, including the continuation
towards the north of the previous current, and the regional
cyclonic circulation within the Angola Dome region;
(iii) The south subtropical Atlantic gyre (15°S-20°S, 40°W-20°W),
corresponding to an important area of warm-salted water
formation by subduction, and feeding the South Equatorial
Current (SEC), which is itself formed by at least three easterly
branches separated by narrow regions of less evident counter-
currents (Stramma, 1991);
(iv) And finally, more to the west, near South America, the zone of
divergence of the southern SEC while approaching the
Brazilian edge.
This last region is in part at the origin of the feeding of numerous current
systems bordering the Brazilian coast, either towards the north or south.
Towards the north, the northern branch of the SEC termination forms the
North Brazilian Current-North Brazilian Under Current (NBC-NBUC) system,
one of the most powerful western boundary currents in the world. This
system participates in the feeding of a few other currents (Stramma et al.,
19
Chapter 1 – Motivation and objective
2005); among them, the northward Guyana Current, the eastward North
Equatorial Counter Current with its associated complex retroflection system
(Goes et al., 2005), and finally the eastward Equatorial Under Current (EUC).
It is believed that NBC accounts for approximately one-third of the net
warm-water transported across the equatorial tropical gyre boundary into
the North Atlantic, compensating for the southward export of NADW
(Stramma, 1991; Stramma and England, 1999) as previously cited. More to
the south, the southern branch of SEC forms the Brazil Current (BC)
propagating southward along the coast of Brazil meeting the Malvinas
Current (Gordon and Greengrove, 1986) at about 35°S, which is itself fed
partially by cold water coming from the Pacific Ocean via the Drake’s
passage. The latitude where the SEC bifurcation occurs is not yet very well
known, although it has been demonstrated that the North Brazil Under
Current (NBUC) originates south of 10°S (Stramma et al., 2003, 2005; Schott
et al., 2002, 2005), where a branch of the SEC flows northward and merges
within the NBUC. The BC, which flows southwesterly, merges into the
Atlantic Meridional Overturning Cell (MOC) system (Silveira et al., 1994;
Stramma et al., 1990, 1995; Schott et al., 1998) in the vicinity of the Brazil-
Malvinas confluence (Olson et al., 1988; Garzoli and Garraffo, 1989).
The interactions between sea surface temperature (SST) and the easterly
atmospheric circulation may play a significant role in local climate
fluctuations of Northeast Brazil, a region affected by intermittence of severe
droughts and floods (Moura and Shukla, 1981; Rao et al., 1993).
In the oligotrophic regions of the ocean, the supply of inorganic nutrients
to the euphotic layer may limit the concentration of microalgal biomass, the
rate of phytoplankton growth, or both (Marañón et al., 2000). Several
Mechanisms have been proposed to explain the formation and maintenance
of the deep chlorophyll maximum (DCM), that is a oceanographic feature of
tropical and subtropical oceans. Higher in-situ growth at the nutricline than
in the upper mixed layer, and acclimation of phytoplankton to low
irradiance, has been suggested (Cullen,1982) as cause of DCM formation at
subtropical gyres.
Very early reports of low productivity in the subtropical gyres
20
Chapter 1 – Motivation and objective
(Steeman-Nielsen and Jansen, 1957; Thomas, 1970; Eppley et al., 1973)
have been considered to be in error due to methodological inadequacies
(Fitzwater et al., 1982). Nevertheless, recent work carried out in oligotrophic
regions, using new techiniques (Malone et al. 1993; Letelier et al. 1996),
presents a large range in the rates of phytoplankton production and growth
inside these areas. In fact, this controversy will be maintained, while the lack
of knowledge about temporal and spatial variability in the biology of open
ocean is being kept (Marañón et al., 2000). Consequently, any uncertainties
in the productivity estimates in these areas will have significant effects on
the prediction of global biogeochemical models (Marañón et al., 2000). In
other hand, there are few studies about the physical processes related with
DCM formation inside the southwestern tropical Atlantic. Montes (2003)
suggests that the support of nutrient inside euphotic zone, in the
southwestern tropical Atlantic, is maintained by diffusion process in
nutricline basis.
Supported by the ideas stated above, the goal of this work is to investigate
the seasonal and intraseasonal variability of the western boundary regime of
the South tropical Atlantic. I will conduct this research following these
pathways:
(a) The use of a numerical model in the Tropical Atlantic to investigate the
seasonal and intraseasonal variability of main currents and SST along
equatorial band.
(b) The investigation of the near surface ocean heat budget in the
Southwestern tropical Atlantic from high resolved modelling results.
(c) The investigation of the zonal wind stress variability as an important
forcing for the intraseasonal variability of the southern branch of SEC
regime.
(d) The influence of sSEC on the distribution of phytoplankton biomass
along the southwestern boundary of the tropical Atlantic.
This manuscript is organized as follows: the next chapter brings an
21
Chapter 1 – Motivation and objective
22
overview of the study area and the used datasets. The main characteristics
of ROMS and conditions of the simulations are presented in Chapter 3.
Chapter 4 shows illustrations of the tropical Atlantic simulation, where,
simulated and observed thermal structures in upper ocean layer are
compared. Also in Chapter 4, intensities of simulated oceanic currents, as
well as the mass transport across three zonal sections along the equator are
presented. Chapter 5 presents simulated and observed thermal structures in
the upper layer to improve the climatological high resolved simulation of
Southwestern tropical Atlantic. Chapter 5 also provides a simulated
seasonal evolution of oceanic components that compound the heat content
in the mixed layer, as well as mass and heat transport across three zonal
sections and a fourth section along the PIRATA-SWE sites. Chapter 6
presents the results of an interannual simulation in the same region of the
Southwestern tropical Atlantic (seen in chapter 5). In Chapter 6 the ROMS
evaluation is performed by comparing outputs of the model with observed
high-frequency SST fields estimated by satellite, thermal structures and
mixed layer depths provided by the PIRATA-SWE dataset. As a first
dynamical application of ROMS, it shows examples of simulated variation of
mass transports across three zonal transects, and a section along the
PIRATA-SWE sites. In Chapter 7, results of biogeochemical modelling of the
oligotrophic zone in the Southwestern tropical Atlantic are presented. The
last Chapter presents the conclusion and perspectives.
Chapter 2 – Observational and derived data
Chapter 2
2. Observational and derived data
This thesis concentrates on the tropical Atlantic with specific focus at the
upper layer of the western boundary of the Southern tropical Atlantic along
the Brazilian shelf. The purpose here is to introduce the data necessary to
feed surface and open boundaries of the modelling package chosen, The
Regional Ocean Model system (ROMS). The data presented in this chapter
can describe the seasonal and intraseasonal mechanisms of the coupling
ocean and atmosphere variability. Thus, results from the ROMS simulation
were obtained to evaluate this modelling tool representing these regions.
23
Chapter 2 – Observational and derived data
2.1. Surface forcing datasets
2.1.1. Comprehensive Ocean-Atmosphere Data Set (COADS)
The Comprehensive Ocean-Atmosphere Data Set (COADS) is a
extensive and widely used set of sea surface data. The COADS is maintained
by U.S. inter-agency cooperation since January of 1981, formed by NOAA’s
Environmental Research Labs. (ERL), Cooperative Institute for Research in
Environmental Sciences (CIRES), NCDC and NCAR. The COADS data is
actually composed by observational data from a variety of ship cruises,
drifting and moored buoys from Canada’s Marine Environmental Data
Service (MEDS) and Pacific Marine Environmental Laboratory (PMEL) and
the National Data Buoy Center (NDBC) (Wooddruff et al., 1998)
The basic observed variables in COADS include sea surface and air
temperatures, wind, humidity (dew point or wet bulb temperature),
barometric pressure, cloudiness weather, and wave and swell fields. The
observations have been quality controlled, and monthly statistics calculated
for the period of 1854 to 1995, for resolution of 2º x 2º boxes and 1º x 1º
from 1960 to 1993 (Wooddruff et al., 1998; daSilva et al., 1994).
2.1.2. QuikSCAT daily wind data
The wind stress data were obtained from the CERSAT/IFREMER
(Centre ERS d’Archivage et de Traitement/Institut Français de Recherche
pour l’Exploitation de la Mer ) available on line at http://www.ifremer.fr/
using OPENDAP tool (http://www.opendap.org
), to be used in interannual
runs. The scatterometer winds are retrieved from Quick Scatterometer
(QuikSCAT) the version 3 (Penven et al., 2008, Liu, 2002). It consists of daily
averaged gridded values of scalar wind speed, meridional and zonal
components of the wind velocity and wind speed squared. The grid has 1.8º
latitude x 1.6º longitude resolution (Tsai et. al., 2000).
24
Chapter 2 – Observational and derived data
2.2. PIRATA data set
Entering now in its permanent phase, PIRATA is an ocean-
meteorological observing system which contributes to the operational climate
survey of the global ocean (Servain et al., 1998; Bourlés et al., 2008). PIRATA
is the result of an international scientific cooperation gathering Brazil,
France and the United States of America. Its main scientific motivation is to
monitor, describe and understand the two main modes of climatic variability
over the tropical Atlantic which are the equatorial and the meridional modes
(Servain et al., 1998). Based on this rational, the original array configuration
is presently composed by 10 ATLAS moorings (Hayes et al., 1991) along the
equatorial line (35°W, 23°W, 10°W and 0°E) and two meridional lines, one
along the 38°W meridian (4°N, 7°N, 12°N and 15°N), and the other along the
10°W meridian (0°N, 6°S and 10°S). Brazil and France are responsible for the
yearly sea maintenance of the array, while USA provides the ATLAS material
and process the raw data. Daily transmissions of the first 500m of ocean
temperature (11 levels) and salinity (4 levels), and the meteorological main
variables at the sea surface, are ensured thanks to the ARGOS satellite
system, and immediately available on the Web after first validation
(http://www.pmel.noaa.gov/ pirata/).
Once assured the success of the PIRATA original array (Bourlés et al.,
2008), and in order to complete the necessary set of observations to
understand more in details the seasonal and inter-annual variability over
specific areas in the tropical Atlantic, it was decided to extend the original
mooring array. Three extensions were proposed: a first one over the
southwest tropical basin (along the Brazilian coast), a second one over the
southeast basin (along the African coast, south of the Equator), and a third
one over the northern and northeast basin (again along the African coast).
Brazil took alone the responsibility of implementing the south-western
extension. Obviously, the main goal of this project, the PIRATA South West
Extension (PIRATA-SWE), is primarily to help in the forecasting of the
Brazilian climate, especially over the Northeast Brazilian region.
Three sites for the ATLAS mooring were chosen for the PIRATA-SWE
(Figure 5.1) according to scientific arguments (Nobre et al., 2005, 2008): (i)
25
Chapter 2 – Observational and derived data
8ºS-30ºW in connection with the 0ºN-10ºW site of the PIRATA’s backbone,
(i.e. the equatorial Atlantic warm pool vs. the cold tongue complex); (ii) 14ºS-
32ºW in connection with the SEC flow (properties and heat flux transport
variability); (iii) 19ºS-34ºW in connection with the SACZ (relations to the
oceanic heat flux changes). Obviously, the PIRATA-SWE array only addresses
the changes in the thermohaline structure within the SEC's flow into the
bifurcation region. The understanding of changes in the partition between
the NBC and the BC at the SEC bifurcation is out of the PIRATA-SWE
objectives.
The three ATLAS buoys of the PIRATA-SWE were launched during the
second half of August 2005, thus the time series analysed here cover the
first year of dataset, from September 2005 to August 2006. With the
exception of a four-month interruption of the 8°S-32°W site due to anchoring
problems (the buoy’s nylon cable sheared off after ten months of
deployment), the buoys reported daily with nearly 100% of data return. In
this study we are using the daily observed PIRATA-SWE data for temperature
(0m, 20m, 40m, 60m, 80m, 100m, 120m, 140m, 180m, 300m, 500m)
operationally obtained on the Web at the PMEL/NOAA address
(http://www.pmel.noaa.gov/pirata/).
2.3. Hydrology and biogeochemical datasets
2.3.1. WOA2005 data
The World Ocean Atlas (WOA2005) is a data product of the Ocean
Climate Laboratory of National Oceanographic Data Center (NODC). The
WOA consists a climatology with global grids at 1º spatial resolution of three-
dimensional fields (Conkright et al., 2002). The data are normally
interpolated onto 33 vertical layers from surface to seafloor. WOA fields
include eight ocean properties (temperature, salinity, dissolved oxygen,
apparent oxygen utilization, percent oxygen saturation, phosphate, silic acid
and nitrate). WOA fields are produced for annual, seasonal and monthly
time-scales. This data product is distributed with ROMSTOLS (Penven et al.,
2008) package version used in this thesis.
26
Chapter 2 – Observational and derived data
2.3.2. ECCO interannual model data
The ECCO Ocean Data Assimilation is a project which study seasonal
to interannual changes in ocean circulation from assimilating observations
with a global circulation model. The model assimilates altimetric data
provide from TOPEX/POSEIDON data, CTD and XBT’s profiles, current
meter, drifters, floats and air sea-flux from NCEP reanalyses. The ECCO
model domain has 46 vertical levels and a telescopic horizontal grid of 1º
covering the globe (80ºN to 80ºS) and 1/3º within 20º of the equator.
Hydrology ECCO results from 2000 to 2007 were obtained through the use
of OPENDAP tool (http://www.opendap.org) to feed the boundary and initial
conditions of interannual run of high resolution ROMS simulation of the
southwestern tropical Atlantic.
2.3.3. SeaWifs data
The SeaWifs chlorophyll a data is a derived product from images from
SeaWifs scanner and algorithms (O’ Reilly et al., 1998). The processed data
can be obtained at different time-scales (monthly, 8-day, daily). It is available
in http://las.pfeg.noaa.gov/OceanWatch/ocean_watch_safari.php
with
horizontal resolution of 0.1º. The Climatological run of biogeochemical model
used seasonal Seawifs data of Chlorophyll a to build boundary and initial
conditions is seasonal and was obtained with ROMSTOOLS (Penven, et al.,
2008).
27
Chapter 2 – Observational and derived data
28
2.4. MLD Criteria
Most criteria used to determine the MLD in the ocean require that the
deviation of the temperature T (or density,
t
σ
) from its surface value be
smaller then a certain fixed value (Sprintall and Tomczak, 1990; Brainerd
and Gregg, 1995). The MLD is estimated as the depth at which density is
equal to the sea surface value plus an increment
t
σ
Δ
equivalent to a desired
net decrease in temperature. For instance, Miller (1976) and Spall (1991) use
()
0125.0
tt
σ=σΔ to determine the mixed layer depth, while Sprintall and
Tomczak (1992) and Ohlmann et al. (1996) adopt
(
T/C5.0
t
o
t
σ=σΔ
)
, where
T/
t
σ is the coefficient of thermal expansion. Following Sprintall and
Tomczak (1992), we evaluate the MLD in terms of temperature and density
steps ( TΔ =0.5°C and
(
)
T/C0.5
tt
σ
°=σΔ ) from the SST and density (
(
)
0T
and
()
0
t
σ ) obtained from the PIRATA and ROMS profiles:
()
Δ
σ
+σ=σ= T
T
0zMLD
t
tt
(A.1)
where
T/
t
σ is calculated as a function of the surface temperature and
salinity (Blanck, 1999).
Chapter 3 – The modelling tools
Chapter 3
3. The modelling tools
To evaluate interactions between ocean and atmosphere, its variability, and
distribution transport, inside a study area, a numerical modelling related to
low operational cost and to the possibility of studying large geographic areas
were chosen. Based on the use of a numerical modelling tool, the focus of
this work is to investigate a seasonal and intraseasonal variability of the
main transport systems, heat budget and biogeochemical cycles. In this
chapter, the Regional Ocean Modeling System (ROMS) is presented, with
more details on definitions of grids, boundary and stability conditions.
29
Chapter 3 – The modelling tools
3.1 The Regional Ocean Modeling System - ROMS
ROMS is an ocean model (Shchepetkin and McWilliams, 2005)
previously adapted to different regions of the world (Haidvogel et al., 2000;
Malanotte-Rizzoli et al., 2000; She and Klinck, 2000; Penven et al., 2000,
2001a,b; McCready and Geyer, 2001; Lutjeharms et al., 2003). The model
solves the free surface primitive equations in an Earth-centered rotating
environment based on the classical Boussinesq approximation and
hydrostatic vertical momentum balance. ROMS is discretized in terrain-
following vertical coordinates. The model grid, forcing, initial and boundary
conditions were built using the ROMSTOOLS package developed by the
Institut de Recherche pour le Développement (IRD) (Penven et al., 2008).
Upstream advection is treated with a third-order scheme that enhances the
solution by generating steep gradients as a function of a given grid size
(Shchepetkin and McWilliams, 1998). Unresolved vertical subgrid-scale
processes are parameterized by an adaptation of the non-local K-profile
planetary boundary layer scheme (Large et al., 1994). A complete description
of the model can be found in Haidvogel et al. (2000) and Shchepetkin and
McWilliams (2005).
3.2 Model Grids
Reproducing ocean features along the Equatorial band required a
horizontal regular grid of 341 x 245 cells with a resolution of 1/6° in an area
that goes from 20ºS to 20ºN and between 42°W and 15°E (Figure 3.1). The
vertical discretization used had 30 sigma levels (terrain-following). For
application in the Southwestern tropical Atlantic, near the NE Brazilian
coast, the integration domain used was from 25°S to 5°S and from 47°W to
20°W (dashed line in Figure. 3.1), with a isotropic 1/12° horizontal grid with
323 x 249 horizontal mesh cells and a vertical discretization of 40 levels.
30
Chapter 3 – The modelling tools
Figure 3.1. Region of model integrations, solid lines indicate domain of 1/6º
climatological equatorial modelling. Dashed lines indicate southwestern tropical
Atlantic domain limits of 1/12º horizontal grid.
At both presented cases, bottom topography was derived from a 2’
resolution database ETOPO2 (Smith and Sandwell, 1997). A slope parameter
20.0<= hhr
was used to prevent errors in the computation of the pressure
gradient (Haidvogel et al., 2000).
3.3 Boundary Conditions
At the open boundaries, an active, implicit, upstream-biased radiation
condition connects the model solution to the surrounding ocean
(Marchesiello et al., 2001). Horizontal Laplacian diffusivity inside the
integration domain is zero, and 1 degree relaxation scheme is imposed (up to
10
4
m
2
s
-1
) in the sponge layers near the open ocean boundaries. The model
equations are subjected to no-slip boundary conditions along the coastline.
31
Chapter 3 – The modelling tools
The oceanic circulation was forced on sea surface by monthly-averaged
winds, heat and fresh water fluxes derived from COADS (daSilva et al., 1994)
for the climatological simulations (Chapters 4, 5 and 7). For the interannual
run (Chapter 6) a daily-mean (2005-2007) wind stress by QuikSCAT (Liu,
2002) replaces the COADS wind dataset. To compose the lateral forcing of
climatological run, hydrography and geostrophic velocities derived from
WOA2005 (Conkright et al., 2002) were used. On the other hand, a basin-
scale daily-mean (2005-2007) hydrography and currents derived from
OGCM–ECCO (http://ecco.jpl.nasa.gov/) was used to infer thermohaline
properties and water volume exchanges at the open horizontal boundaries.
3.3 Biogeochemical model coupled to ROMS
The ROMS version used in this thesis presents the possibility of being
coupled with a biogeochemical model NCPZD (Moore, 2002). This model is
based on limitant nutrient. Its hypothesis considers twelve variables of state:
nitrate, ammonia, Chlorophyll a, phytoplankton, zooplankton, micro and
macro detritus (for nitrogen and carbon), oxygen, inorganic dissolved carbon
and pH. These variables are balanced during each time step according to
biogeochemical reactions and to horizontal and vertical advectives transport,
as well as to vertical diffusion transport calculated by the physical model.
More details about equations can be found in ROMS documentation
(http://roms.mpl.ird.fr/documentation.html).
3.4 Model Simulation
The climatological simulations ran from a state of rest during 10 years.
After a spin-up period of one year, the model had achieved a statistically
steady state. All the numerical results from these runs examined correspond
to the last year of simulation (year 10).
The interannual simulation was performed in two steps. First, the
32
Chapter 3 – The modelling tools
33
rested ocean was gradually forced by applying ramp functions to the
boundary conditions for the period Jan.-Dec. 2001. After this, the full sets of
forcing conditions for 2001-2007 were considered for the interannual
simulation. Only results from 2005-2007 were used.
Chapter 4 – The pathways and thermal structure of surface currents in the…
Chapter 4
4. The pathways and thermal structure
of surface currents in the equatorial
Atlantic Ocean
This chapter aims to use a regional well-solved circulation model to
investigate the subsurface zonal currents and the thermal structure of the
upper 500 m equatorial ocean layer. The simulations were performed with
the Regional Ocean Modeling System (ROMS), which improves the numerical
resolution of strong currents and counter-currents systems like those
observed in equatorial Atlantic.
Nowadays, the knowledge of the pathways and the dynamics of the
oceans through prediction models can contribute to solve many questions on
climate and oceanography sciences. Modeling constitutes an essential tool
for evaluation of the standings and variations presented in a complex
hydrodynamic regime, just like the equatorial Atlantic Ocean. Using the
available records of observed PIRATA data, Foltz et al. (2003) showed that
the latent heat loss and the absorbed shortwave radiation represent the
largest contributions for the seasonal SST variability in the northwest and
southeast tropical Atlantic basin. Over the equator, however, the authors
34
Chapter 4 – The pathways and thermal structure of surface currents in the…
suggest that the contributions from the latent heat loss are less important
than the horizontal temperature advection and the vertical entrainment.
Chaves and Nobre (2004) by observational and numerical modeling results
suggest that atmospheric variability over the southwest tropical Atlantic also
influences the SST variability locally, thus forming a coupled system with
multiple feedback processes.
4.1 The ROMS Approach
The simulation performed here consists in a meso-scale scenario that
reproduced the equatorial currents system. The model domain results in 341
x 245 horizontal mesh cells in a regular horizontal grid with a resolution of
1/6
°
covering an ocean area from latitude 20ºS to 20ºN and from 42ºW to
15ºE of longitude. The vertical discretization used had 30 levels. Bottom
topography was derived from a 2’ resolution database ETOPO2 (Smith and
Sandwell, 1997). At the three open boundaries (north, west and south) an
active, implicit, upstream biased radiation condition connects the model
solution to the surrounding ocean (Marchesiello et al., 2001).
Figure 4.1. Study area and model integration domain. The sections A, B and C are
dashed lines and the squares over the equator line are the PIRATA mooring sites.
Bathymetry is represented by gray shading contour.
35
Chapter 4 – The pathways and thermal structure of surface currents in the…
Horizontal Laplacian diffusivity inside the integration domain is zero,
and a 5-point smooth increasing is imposed (up to 800 m
2
.s
-1
) in sponge
layers near open ocean boundaries. The model domain was forced at the sea
surface by winds, heat and fresh water fluxes derived from the
Comprehensive Ocean-Atmosphere Data Set (COADS) monthly fluxes data at
0.5°x0.5° resolution (da Silva et al., 1994). The database World Ocean Atlas
2001 (WOA2001), monthly climatology at 1°x1° resolution (Conkright et al.,
2002), is used to infer thermodynamics (temperature and salinity) and
geostrophic currents at the open boundaries.
4.2 Experimental dataset
The horizontal distributions of monthly-averaged Sea Surface
Temperature (SST) obtained in simulations were compared to satellite
observations. The Advanced Very High Resolution Radiometer (AVHRR)
aboard the National Oceanic Atmospheric Administration (NOAA) satellite
was employed in comparisons. The data set from AVHRR-SST has a 1/18°
resolution and the observation period was 1985-2006. This dataset is
available on <http://dods.jpl.nasa.gov/>.
4.3 ROMS Simulation Results
Using the forcing boundary conditions presented in previous section
4.1, the model ran from a state of rest during 10 years. Figure 4.2 shows a
time series of the integration domain kinetic energy (KE) from the model. It
indicates that after a spin-up period of two years, the model has achieved a
statistically steady state. After stabilization, the model had only small
oscillations, standing a mean kinetic energy around 1,4 J.m
-3
. These
oscillations represent the seasonal changes in each year. All the numerical
results correspond to the last year of simulation (year 10).
36
Chapter 4 – The pathways and thermal structure of surface currents in the…
Figure 4.2. Mean kinetic energy (J.m
-3
) of model (ROMS) during 10 simulation years.
4.3.1 Current structure and zonal transport
The equatorial currents obtained in numerical simulations were
identified and characterized along the three sections located at 25°W, 15°W
and 5°W (see Figure 4.1). According to Lumpkin and Garzoli (2005), the SEC
system is part of the equatorial gyre. This flow is clearly divided in two
westward jets, the SECc and the SECnc, separated at 0-2°S by a band of
weak southward flow, which is induced by the equatorial divergence.
Numerical results confirm these observations, with the westward SECc flow
occurring between 1°S–8°S, as well as the westward SECn transport near
surface in the range ~1°N–4°N, between the equator and the NECC (Figure
4.3 to 4.7).
Model results show that the eastward flow of NECC occurs at ~4°N–
8°N, with an indistinct core in function to the time of the year (Figure 4.3 to
4.7). These results corroborate with the description of Richardson and
McKee (1984). They found an eastward NECC flow close to surface in
latitudes 3°N–10°N with velocities reaching up to 0.5 m.s
-1
.
37
Chapter 4 – The pathways and thermal structure of surface currents in the…
Figure 4.3. Zonal transport (u-component) obtained from model simulation (ROMS)
along Section A, 25°W. Dashed lines represent the isotherm 20°C. Monthly-averaged
values for March and August.
The EUC was present in simulations below the equator line on all
ocean basin. These results agree with the findings of Stramma and Schott
(1999), which state that the EUC crosses the entire Atlantic, feeding more or
less the equatorial upwelling.
Figure 4.4. Seasonal variability of EUC, NECC, SECc and SECn transports at 25W°
section, obtained in simulations.
38
Chapter 4 – The pathways and thermal structure of surface currents in the…
Numerical results presented in Figure 4.3 show at 25°W (section A,
Figure 4.1) a core of the EUC located closer to sea surface in March (~80 m,
velocity of 0.8 m.s
-1
) than in August (~100m, velocity of 1.0 m.s
-1
). Giarolla et
al. (2005) pointed out that the EUC is shallower during January to May and
deeper in other parts of the year. By the same time, the sea measurements of
Stramma et al (2003), obtained during March-April 2000 in a 23°W section
show a EUC core velocity of 0.75 m.s
-1
at ~80 m depth. Simulation results
are also in good agreement with the recent analysis of Brandt et al. (2006).
These authors measured a EUC core (0.7 m.s
-1
) positioned at 85 m along the
26°W section. The associated EUC transport calculated from model results
also presented a good fit to values estimated from sea measurements. While
Brandt et al. (2006) found 13.8 Sv (1 Sv 10
6
m
3
.s
-1
) in the section 26°W,
ROMS model pointed out 12.3 Sv (March) and 15 Sv (August) at the 25°W
section (Figure 4.3).
The two main equatorial SEC branches were sharply defined in
simulations, separated by the EUC. Numerical velocities of SECc reach -0.5
m.s
-1
and SECn reach -0.6 m.s
-1
in March. In August the SECc and the
SECn were more extended, with velocities reaching -0.2 m.s
-1
. s
Measurements (Brandt et al., 2006) at 26°W furnished westward transports
of -4.9 Sv (SECc) and -3.5 Sv (SECn). Model issues show good agreement
with these estimations, especially for the central SEC.
Simulations showed that the NECC is more defined during March,
with a core velocity of 0.4 m.s
-1
located near 5°N. In August, a deeper current
core was present in this same latitude (Figure 4.3), with velocities of 0.3 m.s
-
1
. According to Richardson and McKee (1984), the NECC responds strongly
to seasonal variations on atmospheric forcing. Around March, when the
Intertropical Convergence Zone (ITCZ) is at its southernmost position, the
equatorial trade-winds are weakest while in August, when the ITCZ is at its
northernmost position, the equatorial winds are strongest (Weisberg, 1985).
The seasonal variability of the transport close to equator due to main
zonal currents at 25°W (section A) is presented in Figure 4.4. In general one
may say that the transport along the 25°W section did not show strong
seasonal variability. EUC presented maximum values during April–August
39
Chapter 4 – The pathways and thermal structure of surface currents in the…
(21.8 Sv), and lowest transport in the period September to March. The NECC
variability shows two maxima in April and July, 10.1 Sv and 10 Sv,
respectively. These numerical results mismatch with Richardson and McKee
(1984) contributions, which found that NECC has peaks of transport at
25°W-30°W profiles during July-September. The equatorial branches of the
SEC (SECc and SECn) presented minimum westward transports between
May and September, which is in contrast to the period of maximum
eastward transport by EUC.
Figure 4.5. Zonal transport (u-component) obtained from model simulation (ROMS)
along Section B, 15°W. Dashed lines represent the isotherm 20°C. Monthly-averaged
values for March and August.
The comparison between mean zonal current structures obtained
numerically for March and August at 15°W (section B) is shown in Figure
4.5. The seasonal variability of the zonal transport at this longitude is
presented in Figure 4.6. At 15°W the SECn core is deeper than the SECc
core, although both westward currents are sharply in March when compared
to the weaker and sparser structures found in August. The SECc reaches
velocities of -0.5 m.s
-1
during March and drops to -0.15 m.s
-1
in August. The
SECn showed velocity core of -0.7 m.s
-1
in March, which is reduced to -0.4
m.s
-1
during August. Figure 4.5 shows that both currents have almost the
same mean net transport in March (-11.8 Sv, SECn) and August (-11.9 Sv,
SECc). During the austral winter (August), weakest westward transports are
verified in simulations, when the SECn drops to -4.1 Sv and SECc to -2.8 Sv
(Figure 4.5 and 4.6).
40
Chapter 4 – The pathways and thermal structure of surface currents in the…
Figure 4.6. Seasonal variability of EUC, NECC, SECc and SECn transports at 15W°
section, obtained in simulations.
Numerical results indicate that the EUC core at 15°W has the same
maximum velocity (0.8 m.s
-1
) in both months of comparison, but it presents
an increase in transport flow that rises from 7.9 Sv in March to 13.4 Sv in
August (Figure 4.5). Simulations also indicate that month-averaged NECC
transport of 5.4 Sv in March is reduced to 2.6 Sv in August. The EUC
presented the highest values of transport in the period April–September. The
SEC branches have the highest values of flow during the months of October
to April, always in phase with the period of reduced values of EUC transport.
The velocity structure, core position and zonal transports at the 5°W
cross-equatorial transect (section C, Figure 4.1) are presented in Figures 4.7
and 4.8. In Figure 4.7 it is possible to identify the Guinea Current (GC), the
equatorial SEC branches and the EUC. The NECC merges with GC near the
Africa Coast and flows into the Guinea Gulf.
Simulations also indicated that although the core velocities reach
almost the same maximum intensity (0.8 m.s
-1
, not shown here) in both
periods, the EUC core is deeper in August (80-110m depth) than in March
(50–70m depth). In contrast, the SEC branches were more defined in March.
During this period the SECc at 5°W is shallower and more extended than
41
Chapter 4 – The pathways and thermal structure of surface currents in the…
SECn, both with cores velocity of -0.5 m.s
-1
. In August, the SECc and SECn
present the weakest values along the last simulated year, with core velocities
around -0.1 m.s
-1
and -0.3 m.s
-1
, respectively. Model results showed more
important GC velocities in March, although the overall transport had been
superior in August (3.8 Sv). In March (1.6 Sv) the GC was pushed against
the continent due to the enlargement of the SECn. Due to reduction of the
current area, in March, the GC gains on velocity, but loses in transport
(Figure 4.7). The simulations show that the mean SECc transport was
sensible reduced in August when compared to March, dropping from -8.2 Sv
to -4 Sv. The contribution of EUC to equatorial mean eastward transport did
not show important differences for both compared periods (March 11.3 Sv
and August 10.3 Sv). As verified in 25°W and 15°W sections, the EUC core is
slightly deeper during August when compared to its position in March
(Figure 4.7).
The seasonal variation of the equatorial transport along the 5°W
section presented in Figure 4.8 shows highest values of the EUC and GC in
the period April–August, and lowest values occurring in the period
September-March. In opposition, the SEC seasonality presents inverse
standards, with weakest SECc and SECn transports in April–August and
strongest westward flows from September to March.
Figure 4.7. Zonal transport (u-component) obtained from model simulation (ROMS)
along Section C, 5°W. Dashed lines represent the isotherm 20°C. Monthly-averaged
values for March and August.
42
Chapter 4 – The pathways and thermal structure of surface currents in the…
Figure 4.8. Seasonal variability of EUC, NECC, SECc and SECn transports at 5W°
section, obtained in simulations.
4.3.2 Thermal structure
The vertical thermal structure of surface layers in Equatorial Atlantic
obtained from ROMS simulations is compared with observations of the
PIRATA Program. Monthly-averaged temperature profiles from these
moorings (1997-2007, 500 m depth) and climatological model results were
used to construct the Hovmüller diagrams presented in Figure 4.9. The
vertical distribution of the simulated temperature fairly agrees with the field
measurements. In subsurface (z > 100 m) the best resemblance (Δ < |0.5°C|)
occurs for the western PIRATA sites (35°W and 23°W). Such weak difference
between model and observed data is noted in March for those sites in the
shallow (0-100 m) and deeper (300-500 m) waters. Larger differences (~ 1 to
3°C), noted between 100 and 300 m for eastern PIRATA sites (10°W and 0°W)
in austral winter continue to be acceptable considering that the ROMS is
running here according to a climatic forcing. Of special interest is the very
well reproduced MLD for the equatorial PIRATA sites, in particular the
43
Chapter 4 – The pathways and thermal structure of surface currents in the…
upwelling induced shallower MLD verified in eastern buoys during austral
winter and late spring (Figure 4.9).
SST results obtained from the climatological run are compared in Fig.
10 with satellite observations for both months, March and August. SST
values provided by the model are close to the satellite observation, especially
taking in consideration that results from ROMS were obtained with 1/6°
horizontal resolution while the AVHRR images are generated with a 1/18°
grid.
Figure 4.9. Comparison of the vertical temperature distribution provided by ROMS
(black dashed contours) with the three PIRATA buoys (colour shading) extending from
surface down to 500 m. Monthly-averaged distributions for the PIRATA period 1997-
2007.
44
Chapter 4 – The pathways and thermal structure of surface currents in the…
45
The warm belt on equatorial region observed in March is well
reproduced numerically, as well as the meridional loss of ocean surface heat
in North/South directions. In particular, one may see the presence of a well
defined pool of warm waters on the southwestern equatorial Atlantic, which
is also observed in August. This so-called Southwestern Atlantic Warm Pool
(SAWP), is marked by high SST values (> 27°C), that extends off the coastline
from the equator to about 12
°
S during austral winter. According to Huang et
al. (1995), the SAWP is one of the regions of warm waters accumulation. In
contrast to these warm waters a Cold Tongue region can be defined as a
region of relatively cold water (< 24°C) that enters the tropical mixed layer
west of 30°W in the Atlantic in late boreal spring in response to intensified
winds and a shallowing of the thermocline (Figure 4.9 and 4.10).
Figure 4.10. Monthly-averaged SST obtained from ROMS model (above) and from
satellite observations (AVHRR-NOAA) (below) for March and August.
Chapter 5 – Near surface transport and heat budget in the Southwestern …
Chapter 5
5. Near surface transport and Heat
budget in the Southwestern Tropical
Atlantic using a high-resolution
numerical modelling system
The Southwestern tropical Atlantic is an interesting ocean area, acting as a
cradle of multiple oceanic-weather forcings of great importance. In this
region, part of the South Equatorial Current (SEC) feeds many western
boundary currents along the eastern Brazilian edge, thus contributing to the
climatic variability over Northeast Brazil. In this chapter, thermal structures,
heat budget in the surface mixing layer and mass transports were obtained
from climatological ROMS ocean simulations, and analyzed in connection
with scarce available long-term observations of PIRATA-SWE dataset.
46
Chapter 5 – Near surface transport and heat budget in the Southwestern …
5.1 The near surface current system along the area
The South Equatorial Current (SEC) divergence occurs in the
Southwestern tropical Atlantic, producing a northern and a southern branch
along the Brazilian coastline. Towards the north, the northern branch of the
SEC termination forms the North Brazil Under Current-North Brazil Current
(NBUC-NBC) system (Silveira et al., 1994; Stramma et al., 1995; Rodrigues
et al., 2007), one of the most powerful western boundary current in the
world. This is the preferential way of connecting the subduction regions of
the subtropical South Atlantic, the eastern equatorial and off-equatorial
undercurrents, as part of the Atlantic Subtropical Cell (STC), which in turn
feeds the equatorial upwelling systems Malanote-Rizolli et al, 2000; Schott et
al., 2005). This region is also an important highway for the Atlantic
Meridional Overturning Circulation (MOC), where the southward flow of
deeper North Atlantic Deep Water (NADW) is compensated by the northward
transfer of near surface warm and intermediate waters, as well as by the
Antarctic Bottom Water (Lumpkin & Speer, 2003; Ganachaud, 2003).
After the bifurcation close to the Brazilian shelf, the SEC also supplies
the Brazil Current (BC) propagating southward along the coast of Brazil
(Stramma, 1991; Peterson & Stramma, 1991; Stramma et al., 1995). Being
in the region of the southeast trade winds and the South Atlantic
Convergence Zone (SACZ), interactions between sea surface temperatures
(SST) and the easterly atmospheric circulation may play a significant role in
local climate fluctuations of Northeast Brazil, a region affected by
intermittence of severe droughts or floods (Moura & Shukla, 1981; Rao et al.,
1993).
5.2 The ROMS approach
The study case presented here involves the open ocean area near the
Brazilian coast. The integration domain is comprised between 5
°
S and 25
°
S,
20
°
W and 47
°
W (Figure 5.1). An isotropic 1/12
°
horizontal grid results in 323
x 249 horizontal mesh cells. Vertical discretization has 40 levels. Bottom
topography was derived from a 2’ resolution database ETOPO2 (Smith &
47
Chapter 5 – Near surface transport and heat budget in the Southwestern …
Sandwell, 1997), and a slope parameter
20.0hhr
<
=
is used to prevent
errors in the computation of pressure gradient (Haidvogel et al., 2000).
Figure 5.1. Model domain (dashed lines) with the PIRATA-SWE locations (filled
triangles) and three sites located along the western boundary region closer to the
Brazilian edge (filled circles). Near shore 100m and 1000m isobaths are plotted.
In three open boundaries (north, east and south) an active, implicit,
upstream biased radiation condition connects the model solution to the
surrounding ocean (Marchesiello et al., 2001). Horizontal Laplacian
diffusivity inside the integration domain is zero, and a 12-point smooth
increasing is imposed (up to 10
4
m
2
.s
-1
) in sponge layers near open ocean
boundaries. The model equations are subjected to no-slip boundary
conditions along the coastline. A basin scale seasonal hydrology derived from
World Ocean Atlas 2005 WOA2005 database (monthly climatology at 1
°
x1°
resolution) (Conkright et al., 2002) is used to infer thermodynamics
(temperature and salinity) and geostrophic currents in the open boundaries.
The oceanic circulation was forced at the sea surface by winds, heat and
fresh water fluxes derived from the Comprehensive Ocean-Atmosphere Data
48
Chapter 5 – Near surface transport and heat budget in the Southwestern …
Set (COADS) monthly fluxes data at 0.5°x0.5° resolution (daSilva et al.,
1994). The model runs from a state of rest for 10 years. The stability is
achieved after a spin-up period of about one year. All the numerical results
examined in this chapter correspond to averages for the last two years of
simulation, except for the snapshots of high-frequency simulations (Section
5.3.1).
5.3 High-resolved climatological ROMS results
Due to its refined spatial resolution (1/12° in latitude and longitude;
40 vertical levels), I should expect the ROMS to resolve the meso-scale
dynamics in the region, which isn’t sufficiently allowed when OGCMs is
used. Thus, as a preliminary test (Section 5.3.1), the high-resolution spatial
model accuracy was evaluated by comparing two extreme seasonal model
SST snapshots with examples of high-frequency SST patterns provided by
satellite measurements. A second type of verification (Section 5.3.2)
presented, was to validate the simulated annual cycle of temperature and
heat content in the mixing layer depth (MLD). The PIRATA-SWE observations
were then used for a rough local thermal evaluation of the model within the
first 500m depth. As an early application of ROMS, a third type of interest
(Section 5.3.3) is to use the simulated results in order to give insight to our
knowledge of the seasonal variation of the main oceanic dynamics
components which are at the origin of the annual SST change in that area.
Finally, a fourth type of interest (Section 5.3.4) is to look at the yearly
averaged oceanic transport simulated by ROMS across zonal sections at the
three PIRATA buoy latitudes, as well as a fourth section along these buoys
locations.
49
Chapter 5 – Near surface transport and heat budget in the Southwestern …
5.3.1 Instantaneous SST evaluation
The SST patterns for the two mid-month September and March, i.e.
during two extreme seasonal conditions, is presented as the first
illustrations of the model simulation. These figures show mainly a seasonal
meridional change in SST along the South Atlantic western boundary. The
so-called South Atlantic Warm Pool (SAWP), marked by high SST values (>
27°C), extends off the coastline from the equator to about 12
°
S during
austral winter (e.g. September, Figure 5.2a) which only includes the region of
the northern PIRATA buoy. Six months later, during austral summer (e.g.
March, Figure. 5.2b), these high SST values invade the whole studied zone,
and warmer waters now bathe the three PIRATA buoys. In connection with
seasonal meridional extension, the SAWP pattern records seasonal zonal
changes, with a more alongshore location during summer. The seasonal
colder waters (< 22°C) extension at the southern limit of the study domain
follows the same meridional progression as the SAWP. These cold waters
cross the southern limit of the studied domain during March. However, along
the coastline, they remain stationary south of 16°S during all the year (Ikeda
& Stevenson, 1978; Carbonel, 2003). It is also noted, as illustrated by these
figures, that ROMS succeeds in resolving meso-scale dynamical processes
such as frontal structures or filaments.
Figure 5.2. Horizontal distribution of SST obtained from ROMS in mid-September (a)
and mid March (b). PIRATA-SWE locations (filled triangles) and three sites located
along the western boundary region closer to the Brazilian edge (filled circles).
Recife
Salvador
Rio de Janeiro
45
o
W 40
o
W 35
o
W 30
o
W 25
o
W 20
o
W
24
o
S
20
o
S
16
o
S
12
o
S
8
o
S
SST - ROMS12 (
o
C) - 09/15
08S-30W
14S-32W
19S-34W
08S-34W
14S-38W
19S-38W
20 22 24 26 28 30
(a)
Recife
Salvador
Rio de Janeiro
45
o
W 40
o
W 35
o
W 30
o
W 25
o
W 20
o
W
24
o
S
20
o
S
16
o
S
12
o
S
8
o
S
08S-30W
14S-32W
19S-34W
08S-34W
14S-38W
19S-38W
SST - ROMS12 (
o
C) - 03/15
(b)
20 22 24 26 28 30
50
Chapter 5 – Near surface transport and heat budget in the Southwestern …
These phenomena, which are particularly visible at the frontier
between warm and cold waters on the ROMS results (Figures. 5.2a,b) when
using a 1/12° resolution (= 9.25 km), also appear (slightly more diffuse) in
two examples of daily observed SST by satellite also using 9,25 km
resolution (GODAE High Resolution Sea Surface Temperature Pilot Project -
GHRSST-PP, 2007): one in September 15, 2005 (Figures 5.3a), the other one
in March 15, 2006 (Figure 5.3b). Due to several individual images used to
estimate the satellite SST (process brought about cloud covers) the
mesoscale structures seem damped when compared to the model
instantaneous SST outputs.
(b)
(a)
Figure 5.3. Horizontal distribution of SST obtained from observed satellite data on 15
September 2005 (a) and 15 March 2006 (b). PIRATA-SWE locations (filled triangles)
and three sites located along the western boundary region closer to the Brazilian edge
(filled circles).
5.3.2 Seasonal evaluation of the heat content
Figures 5.4a,b,c provide the seasonal variations of the 0-500m
temperature for the three PIRATA-SWE locations (08°S-30°W, 14°S-32°W
and 19°S-34°W). These Hovmüller diagrams show the ROMS simulated
temperature (dashed contours), superimposed by daily observed in-situ data
(shaded colors) during the initial 2.5 year, approximately, recording of
PIRATA-SWE network (September 2005 to February 2008). No-shaded parts
51
Chapter 5 – Near surface transport and heat budget in the Southwestern …
mean missing PIRATA values. A reasonable agreement between ROMS and
PIRATA observations is noted, definitely more satisfactory than using an
OGCM, even when forced by inter-annual surface fluxes (e.g. Nobre et al.,
2008). ROMS reproduces the tightening of the thermocline for the most
equatorial PIRATA site (08°S-30°W). The opposite situation, the relaxation of
vertical gradient, occurs for the two southern sites.
Also of great interest is the apparent capacity of ROMS to simulate
intraseasonal variations of the thermocline depth, which clearly occurs for
the central and southernmost locations on the in-situ data within a 3-4
month periodicity. This is particularly evident on the 19°S-34°W site (Figure
5.4c). These observed variabilities can be partially explained in terms of
ocean adjustment to disturbances in the buoyancy field due to the
propagation of barotropic Kelvin waves. It causes vertical displacement of
isopycnals and propagates with the coastal boundary on the left in the
southern hemisphere. Keeping in mind that the present run is only forced by
climatological fluxes, let us consider a promising accuracy of the intra-
seasonal phenomena if ROMS is forced into inter-annual mode (showed in
Chapter 6).
52
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(c)
(b)
(a)
Figure 5.4. Comparison of the monthly mean vertical temperature distribution (ºC)
provided by ROMS (black dashed contours) with monthly mean for the period of
September 2005 to February 2008 of the three PIRATA-SWE buoys (colour shading
and white lines) extending from surface down to 500m. The four vertical black thin
lines correspond to September 15
th
, December 15
th
, March 15
th
and June 15
th
,. (a) 8°S-
30°W; (b) 14°S-32°W; (c) 19°S-34°W.
53
Chapter 5 – Near surface transport and heat budget in the Southwestern …
Detailing the preceding discussion, monthly mean temperature profiles
(0-500m) observed by PIRATA-SWE array during September, December,
March and June (i.e. as related on Figures 5.4a,b,c) are compared to ROMS
outputs for the same calendar months (Figures 5.5a,b,c,d). As shown
previously, and though climatologically forced, the vertical distribution of the
simulated temperature fairly agrees with the field measurements. In
subsurface (z >100m) the best resemblance (Δ < |0.5°C|) occurs for the
PIRATA –SWE northernmost site (8°S-30°W) in March.
Such a small difference between model and observation is also noted,
in December, for that site in the shallow (0-100m) and deeper (300-500m)
waters, as well as for the 14ºS-32ºW site in the same month (December, Fig
5.5b). Larger differences (~ 1-to-2°C), noted below 200 m for the southern
site (19ºS-34ºW) in March, for the two other sites for all months plotted, as
well as upper 100m to 8°S-30°W in June, only, continue to be very
acceptable considering that the ROMS here is running according to a
climatological forcing. Of special interest is the very well reproduced MLD for
the three PIRATA sites, whether during warm season (~20-30m in March), or
cold season (~80-100m in September). SST values provided by the model are
close to the PIRATA observations, except in September for the southern
location, obviously subjected to larger inter-annual variability (see also Fig.
5.3a).
54
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(b)
(a)
(c) (d)
Figure 5.5. September (a), December (b), March (c) and June (d) monthly averaged
vertical profiles of temperature 0-500m at the three PIRATA-SWE sites provided by
climatic ROMS and by PIRATA monthly averaged profiles from September 2005 to
February 2008
55
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(a)
(b)
(Continue)
56
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(c)
Figure 5.6. Comparison of temporal evolution of seasonal heat content in the mixed
layer at 8°S-30°W (a), 14°S-32°W (b) and 19°S-34°W (c), provided by PIRATA in-situ
observations (dashed line) and ROMS (solid line) for the period of September to
August.
Seasonal evolutions of the heat content integrated inside the mixing
layer (MLD defined in Appendix) are computed for the three PIRATA-SWE
sites (Figure 5.6) according to the ROMS outputs (solid lines), and the in-situ
PIRATA measurements from September 2005 to February 2008 (dashed
lines). Like the previous discussion about the thermal profile comparisons
(Figure 5.5), the agreement between simulation and observation here is also
very satisfactory, especially for the site 14°S-32°W, i.e. the centre of the
PIRATA-SWE array, where both observed and simulated curves take very
similar shapes. The two largest differences (~ 4 x 10
8
J.m
-2
), both indicating
a greater observed value of heat content vs. the climatological simulation,
are noted from February to June for the southern site, and from April to
June for the northern site.
57
Chapter 5 – Near surface transport and heat budget in the Southwestern …
5.3.3 Oceanic dynamics vs. atmospheric forcing
After checking that ROMS approach is capable of reproducing the
ocean thermodynamics in the study region, ROMS results was used to test
the relative influence of the ocean dynamics vs. the atmospheric forcing for
settling the seasonal variation of the mixing layer temperature. For that, the
three PIRATA sites are now focused, adding three other locations with the
same latitudes as those of PIRATA sites, but along the Brazilian coast (see
Figure 5.1).
Considering the local heat budget inside the mixing layer, the different
components of the temperature equation can be written as:
{
(
)
{
0
x
'T'u
x
T
U
t
T
FLUXHEAT
T
)DIFFUSIONADVECTION(OCEAN
j
i
j
j
CHANGELOCAL
=Φ
+
+
+
444344421
(5.1)
where U
i
is the mean velocity, T (or SST) is the temperature of MLD, and
T
Φ
is the total heat flux used to force the ROMS (Marchesiello et al., 2003).
Temporal changes of the three terms on the left of Eq. 1 (in °C/Month),
as well as T (in °C), have been monthly integrated through MLD, and are
presented in Figure 5.7 for the three sites of PIRATA-SWE and the three
coastal locations. A first overview indicates that the main seasonal behaviors
of the dynamics are very similar in the whole domain, though we note
somewhat larger amplitudes in higher latitudes than in equatorial areas, and
along the coastline rather than in the open ocean. Highest T (dotted black
curves) values are centered around March (from 28.50°C in the northern
locations, to 27.75°C in the southern locations), and coldest T values are
centered around August-September (from 26.25°C in the north, to 24.00°C
in the south). The seasonal change of temperature is roughly of sinusoidal
type, except in the southern coastal site (Figure 5.7e), located at the eastern
limit of the coldest waters (see Figures 5.2a,b), where T undergoes a rapid
reduction of half degree in November-December, thus slowing down the
58
Chapter 5 – Near surface transport and heat budget in the Southwestern …
thermal increase between minimum (~ 24.8°C) in July-August and maximum
(~ 28.0°C) in February-March. As related on Eq. 5.1 the local changes in T
(continuous black curves on Figure 5.7) results from the combination of net
atmospheric forcing (“FLUX” in red) and net oceanic balance (“ADV+DIFF” in
blue). In agreement with previous works (e.g. Carton & Zhou, 1997), and
except for the location 19°S-38°W, which shall be discussed hereafter, local
variations of T are mainly driven by the atmospheric forcing in the sense that
positive (negative) values of
t
T
generally occur during positive (negative)
values of “FLUX”. It is however interesting to note that, as opposed to what
was generally thought until now for this region, the net oceanic contribution
evolves according to an amplitude of the same order of magnitude (until
3°C/Month, peak-to-peak) rather than the seasonal variations of
atmospheric forcing.
The time duration of positive values of “FLUX”, occurring mainly from
August-September to March-April, (i.e. when the South Hemisphere is
warmer), is significantly longer than the duration of negative occurrences in
May-June-July, i.e. when the sun is at its northern position. These time
durations are slightly modulated according to latitude: close to equator the
“FLUX” is positive during a longer period (~ 8-9 months), while in the south
of the study domain, positive and negative periods tend to be of equal
duration (~ 6 months). Another characteristic is that the positive values of
“FLUX” are generally of larger amplitude than the negative values.
Consequently, in order to maintain the local equilibrium of
t
T
(a longer
duration associated with larger amplitudes) positive “FLUX” must thus be
compensated by an inverse net oceanic contribution during the same
periods. In fact, “ADV+DIFF” opposes to atmospheric forcing “FLUX” most of
the time, with larger and longer negative values during the period August-
September to March-April, and shorter and weaker positive values during
the remainder of the year, centered around June. Additional analyses carried
out from individual oceanic components (not shown here), and confirmed by
results from another recent study (Servain & Lazar, personal
59
Chapter 5 – Near surface transport and heat budget in the Southwestern …
communication), brings some additional information: the cooling of the
mixing layer by oceanic effect noted here in the whole region comes primarily
from a mixing by vertical diffusion between MLD water and deeper colder
waters, while the warming by oceanic effect mainly comes from horizontal
advection and lateral diffusion.
Let us take a closer look at the 19°S-38°W site, in the southern
studied region. It is positioned in a complex region, where the interaction
between southward Brazil Current and Vitoria-Trindade ridge, during the
austral summer, may induce the formation of cyclonic thermocline eddies,
which trap cold waters from an extending upwelling regime north of Cabo
Frio (Castelão et al., 2004; Campos, 2006).
60
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(a)
(b)
(c) (d)
(e)
(f)
Figure 5.7. Seasonal evolution of SST (°C) (black dashed line), atmospheric (red line)
and oceanic (blue line) contributions to the local change of SST (°C/Month) (black
continuous line) provided by ROMS at 8°S-34°W (a), 8°S-30°W (b), 14°S-38°W (c), 14°S-
32°W (d), 19°S-38°W (e) and 19°S-34°W (f) locations.
61
Chapter 5 – Near surface transport and heat budget in the Southwestern …
5.3.4 Mass transports across sections
Mean (yearly averaged) meridional current and transport values
obtained from the ROMS simulation along zonal sections (0-1500m) at 8°S,
14°S and 19°S (i.e. the latitudes of the three PIRATA-SWE buoys) are
presented in Figure 5.8a,b,c. A fourth section (Figure 5.8d) provides current
and transport information within the 0-600m across this mooring track.
Three density levels are indicated on the plots. The
t
σ = 24.5 kg.m
-3
separates the upper Tropical Surface Water (TSW) from the upper
thermocline waters. The
t
σ = 26.8 kg.m
-3
(at about 150 m depth) is the lower
level of the water supplying the EUC (Schott et al., 1998). The
t
σ = 32.15
kg.m
-3
(at about 1100 m depth) indicates the lower boundary of the upper
warmer waters as well as the lower boundary of the Antarctic Intermediate
Water (AAIW). Below 1100 m depth we found the North Atlantic Deep Water
(NADW) extending to about 4000 m depth (
t
σ
= 45.9 kg.m
-3
). This is the
layer where the Deep Western Boundary Current (DWBC) transports
southward cold waters from the North Hemisphere.
The zonal sections (Figures 5.8a,b,c) show a clear representation of the
NBUC-NBC skirting the coastline from 100 to 1000m, with a northward
transport increasing from 5.8 Sv at 19°S to 17.4 Sv at 8°S. The site 8°S-
34°W is located above the core of the NBUC which is situated at about 50
km from the coast, at 180-to-500 m depth approximately (Silveira et al.,
1994; Schott et al., 2002; Stramma et al., 2003; Schott et al., 2005). These
results have been detected from field measurements. Stramma & England
(1999) showed from hydrographic data that the SEC bifurcation latitude is
16°S in near surface layer (top 100m), 20°S in the layer (100-500m) and
26°S in the AAIW (500-1200m). Wienders et al. (2000) also used
hydrographic data to indicate that SEC bifurcation latitude is 14°S at the
surface, 24°S in the (400-500m) layer and around 26°S-28°S in the AAIW.
62
Chapter 5 – Near surface transport and heat budget in the Southwestern …
(a) (b)
(c)
(d)
Figure 5.8. Mean annual volume transport averaged across three zonal sections at
8°S (a), 14°S (b) and 19°S (c), and across the section along the PIRATA-SWE array (d).
Positive (negative) values indicated by solid (dashed) white lines correspond to
northward (southward) currents (panels 8a, 8b, and 8c), while positive (negative)
values indicated by solid (dashed) white lines for the section along the PIRATA array
correspond to eastward (westward) currents (panel 8d). The three horizontal solid
black lines indicate the 24.5, 26.8 and 32.15 sigma-t values (in kg.m-3).
The ROMS results also agree with the numerical findings of Harper
(2000), Malanotte-Rizzoli et al. (2000) and Rodrigues et al. (2007), indicating
a poleward depth increasing of the SEC bifurcation along the Brazilian edge,
as well as its seasonal variability: the SEC bifurcation reaches its
southernmost position in July and its northernmost position in November
(not shown here).
The flowing southward BC, confined to the shallow and near shore
part of the Brazilian continental slope, is especially recognizable on the 19°S
section (Figure 5.8c). These numerical results were observed in previous
63
Chapter 5 – Near surface transport and heat budget in the Southwestern …
64
studies. Miranda & Castro (1981) identified the BC at 19°S as a surface
narrow current (~ 75 km) limited to upper 500 m depth. Evans et al. (1983)
indicate that BC is still confined and organized above the continental shelf at
20.5°S. The World Ocean Circulation Experiment (WOCE) current-moorings
measurements obtained at 19°S pointed out a BC confined to the upper 200
m depth, with a mean southward velocity of approximately 15 cm.s
-1
(Müller
et al., 1998).
The current and transport section along the PIRATA-SWE array (Figure
5.8d) allows us to have an idea of how the southern part of the SEC extends
before reaching the western continental boundary. Indeed, we may note a
succession of more powerful westward systems (one core with 4.8 Sv, from
50 to 200 m depth, and with a 300-400 km width) and less powerful
eastward systems (each one with a core less than 1.0 Sv, from 0 to 100 m
depth, and with a width lower than 100 km). Let us note however that the
most important westward transport of 7.8 Sv shallower 600 m between 19°S-
34°W and 14°S-32°W PIRATA sites has large extension with small velocity
intensity values across the section. Elsewhere in the deeper ocean the E-W
transport given by the model results is relatively weak.
According to ROMS simulation, the annual-mean westward transport
between 8°S-30°W and 19°S-34°W for the SEC is 15.6 Sv for the upper 600
m, which is in good agreement with observations and previous numerical
studies. For instance, Wienders et al. (2000) estimated a westward transport
of 21.5 Sv for the SEC within the SACW layer between 7.5°S to 20°S, while
the hydrographic data of Stramma (1991) showed a westward SEC transport
of about 20 Sv for this same region. Furthermore the recent numerical
simulations of Rodrigues et al. (2007) indicated an annual-mean westward
SEC transport between 6°S and 22°S (along the meridian 30°W) of 18 Sv.
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Chapter 6
6. Interannual high-resolution
regional ocean dynamics simulation in
the Southwestern Tropical Atlantic
Inside this area there is a complex current system of great importance from
an oceanographic point of view, being a key driver for regional climate. A
high resolution ocean model has been adapted for the simulation of this
system, and it is now forced with interannual data sets to be compared with
PIRATA observations. It should be helpful in obtaining a better knowledge of
this oceanic system, its dynamic equilibriums and evolutions. The goal here
is to analyse the accuracy of a versatile new generation state-of-the-art
Regional Ocean Modeling System (ROMS) to reproduce some aspects of
intra-seasonal to interannual ocean dynamics and heat contents in the
region of the southernmost extension of the westward SEC.
65
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
6.1. The high-resolved interannual ROMS approach
The case study presented here involves the open ocean area near the
NE Brazilian coast. The integration domain, bathymetry and boundary
conditions are imposed as presented in the previous chapter. Changes were
made in assimilation of geostrophic currents induced on lateral boundaries,
based on a daily-mean (2005-2007) hydrography scale and currents derived
from OGCM–ECCO (http://ecco.jpl.nasa.gov/
). At sea surface these changes
were based on daily-mean (2005-2007) wind stress by QuickSCAT (Liu,
2002), combined with monthly-averaged heat and fresh water fluxes derived
from COADS (da Silva et al., 1994).
6.2 Thermal structure in Southwestern tropical Atlantic
6.2.1. Instantaneous SST evaluation of interannual simulation
After establishing kinetic energy, in order to validate the numerical
results, the model accuracy was evaluated by comparing daily SST surface
maps derived from the ROMS model to daily infrared SSTs derived from the
GODAE High Resolution SST Pilot Project (GHRSST-PP, 2007).
This comparison was done for two different scenarios: September 15
th
2005, which corresponds to the austral winter time, and March 15
th
2006,
which corresponds to the austral summer time (Figure 6.1). The satellite-
derived SST maps have a horizontal resolution of 9.25 km. A full description
of the GHRSST-PP can be found on the GHRSST-PP web site at
http://www.ghrsst-pp.org.
66
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
(a)
(b)
(c)
(d)
Figure 6.1. Comparison between ROMS-derived SST (a and c) and GHRSST-PP-
derived SST (b and d) for September 15
th
2005 and March 15
th
2006.
In the region of interest, the model and daily SST satellite patterns are
consistent, showing a distinct meridional seasonal change along the South
Atlantic western boundary. The so-called Southwestern Atlantic Warm Pool
(SAWP) (Huang et al., 1995; Nobre et al., 2005), which is marked by high
SST values (> 27°C), extends off the South American coastline from the
equator to about 12
°
S on September 15
th
, 2005, including the northern part
of the PIRATA-SWE (Figures 6.1a, b). Six months later, on March 15
th
, 2006,
these high SST values invade the whole area of study, and warmer waters
are observed at the three PIRATA buoy sites (Figures. 6.1c, d). Besides its
67
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
meridional seasonal migration, the SAWP pattern also records seasonal
zonal changes with a more eastward extension during the summer,
especially in the northern part. The seasonal spreading out of cold waters (<
22°C) in the open ocean in the southern limit of the study domain follows the
same meridional progression as that observed for the SAWP. These cold
waters are advected off the southern limit of the study area in March.
Besides the fact that the model accurately reproduces the satellite-
derived SST overall, it also resolves the mesoscale dynamical processes quite
well, including frontal structures, meanders and local upwelling regions. For
instance, the cold water filaments observed in the model outputs for March
15
th
, 2006 in the vicinity of the Abrolhos Bank (Lat. 17-18
o
S – Long. 38.5
o
’-
39.5
o
W) and the Vitória-Trindade Ridge (Lat. 20°S – Long. 34-38°W) are good
indicators of observed and modeled upwelled waters and mesoscale cyclonic
structures previously documented in this area (Schmid et al., 1995; Campos,
2006). Indeed, the modeled SST is especially efficient compared to the
satellite-derived SST estimates along the coastline where the infrared
retrievals may be cloud-contaminated. Although the horizontal resolution of
the model and the GHRSST-PP SST product are almost the same (~10 km),
the fine mesoscale surface structures evident in the model results seem to be
smoothed in the infrared SST maps. This is probably due to the fact that
several individual images and infrared products are combined and used to
estimate the SST. For example, it does not seem random that higher
differences between satellite and modeled SST are present in late austral
summer close to the coastal region between 16°S and 22°S (Figures 6.1c, d);
this is a well-known cloudy area under the influence of the atmospheric
South Atlantic Convergence Zone – SACZ (Chaves and Nobre, 2004; De
Almeida et al., 2007).
68
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
6.2.2. Seasonal (2005-2007) evaluation of the subsurface temperature
and mixed layer depth
In this second set of model evaluations, samples of local time series of
the vertical distribution of temperature given by ROMS are directly compared
to the first available two-year period dataset provided by PIRATA-SWE.
Figure 6.2a,b,c primarily shows the simulated and observed variations of the
0-500 m temperature profile (Sep 2005-Jun 2007) for the three PIRATA-SWE
locations (08°S-30°W, 14°S-32°W and 19°S-34°W). The modeled temperature
is represented by black contours (for a two-day period), while the PIRATA
temperature appears as white contours (also for a two-day period), and
shaded colors (daily). Vertical interpolations were processed on the PIRATA
data when only one out of three consecutive temperature sensors along the
moored line was dysfunctional. That was the case for the first 50 m level at
the 08°S-30°W mooring site from the 10
th
of November 2006 to the end of
June 2007, as well as for the levels below 180 m at the 19°S-34°W mooring
from the 3rd of November 2006 to the end of June 2007. Missing values are
represented by blank areas. ROMS reproduces well the tightening of the
thermocline for most of the northernmost PIRATA-SWE site (08°S-30°W,
Figure 6.2a), and the relaxation of the vertical gradient for the two
southernmost sites (Figure 6.2b, c). In terms of the seasonal cycle within the
mixed layer, the simulated and observed cores of the warmest (>28°C for the
near equatorial site, >26°C for the two other sites) shallow waters (0-75 m)
occur within the same period of the year. There are, however, a few episodic
discrepancies between ROMS and the PIRATA-SWE data in the levels
including the thermocline and the layer depths from ~75 to ~200 m for the
northernmost site, and from ~100 to ~250 m or even more for the other two
sites. For most of these differences, the 26°-to-16°C isotherm depths given
by the model are shallower than the PIRATA-SWE observations, extending
from a few meters (e.g., the northern site) to a few tens of meters (e.g., the
20°C isotherm depth for a few episodes in central and southern sites). This
indicates a relatively permanent cold bias in the model for these depths,
which can vary from 0.8°C at 08°S-30°W to more than 2°C at 19°S-34°W at
the most. These discrepancies are probably due to the model’s difficulty in
69
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
accurately reproducing the ventilation of subducted lower-thermocline
waters coming from the subtropical South Atlantic Ocean. For the lower
levels (e.g., ~16°-to-12°C isotherms) the model again becomes generally
consistent with the observations.
The PIRATA-SWE data clearly show intraseasonal variations of the
thermocline depth within a 3-4 month periodicity, especially for the central
and southernmost locations (Figure. 6.2b, c). This observed variability can
be partially explained in terms of an ocean adjustment to disturbances in
the buoyancy field due to the propagation of Tropical/Kelvin-Helmholtz
instability waves (Proehl, 1996; Polito and Cornillon, 1997; Jochum et al.,
2004). These disturbances cause vertical displacement of the isotherms and
propagate equatorward keeping the coastal boundary to the left in the
southern hemisphere. Of great interest here is the apparent capacity of
ROMS to generate such variability in the temperature profile (see the black
lines on Figure 6.2b,c around the 20°C isotherm depth), even if simulated
and observed fluctuations are not always in phase.
70
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
(a)
(b)
(c)
Figure 6.2. Comparison between simulated and observed temperature (
o
C) variations
for the upper 500m from September 2005 to June 2007 for the three PIRATA-SWE
locations at (a) 08°S-30°W, (b) 14°S-32°W and (c) 19°S-34°W. Model-derived
temperature is represented by black contours and the PIRATA-SWE observed
temperature appears as white contours and shaded colors.
71
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
In order to gain better insight into the preceding discussion, four
selected monthly mean temperature profiles (0-500 m), computed from
ROMS outputs, are compared to the observational data derived from the
three PIRATA-SWE moorings for the same months (Figure 6.3), i.e.
September 2005, December 2005, March 2006 and June 2006. For most of
the 12 temperature profiles (also marked by vertical black lines on Figure
6.2), the model and observations are in good agreement. This is especially
the case during the first two months (September and December 2005), where
a cold bias, such as those previously discussed, occurs in September 2005
at 14°S-32°W within 100-250 m, and in December 2005 at 19°S-34°W within
50-300 m. Other examples of limited negative biases are observed in March
and June 2006 at these same depths.
We conclude this subsection with an evaluation of the temporal
evolution of the mixed layer depth (MLD) for both the model runs and the
observations. Following Sprintall and Tomczak (1992), we computed the
MLD in terms of temperature and density steps (see the Appendix for
details). Note that for these calculations, the number of available levels for
the modelled temperature and salinity is 20 for the first 300 meters, i.e. two
(five) times the temperature (salinity) number of levels of the PIRATA-SWE
data for the same depth. Figure 6.4 shows the simulated and observed
temporal evolutions of the MLD from September 2005 to July 2007
according to the ROMS outputs (solid lines) and the in-situ PIRATA-SWE
measurements (dashed lines). The ROMS-derived MLD vs. the PIRATA-SWE-
derived MLD shows a similar pattern between the model and the
observations with a few discrepancies that are more pronounced at the 8
o
S-
30
o
W mooring site.
72
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure 6.3. Comparison of monthly averaged vertical profiles of temperature for the
first 500m between the PIRATA-SWE in-situ observations (dashed line) and the ROMS
simulation (solid line) for: (a) September 2005, (b) December 2005, (c) March 2006 and
(d) June 2006. Colours are associated to each PIRATA-SWE buoy: 08°S-30°W (blue),
14°S-32°W (black) and 19°S-34°W (red).
The calculated cross-correlation coefficients between observed and
modelled MLD signals are 0.65 (8°S-30°W), 0.45 (14°S-32°W) and 0.71 (19°S-
34°W) (95% confidence level and 321 degrees of freedom). The cross-
correlation coefficient is a measurement of similarity between the signals.
Quite good agreement is evident and the observed and simulated MLD
curves have similar shapes, ranging from small values (~30-60 m) during the
austral summer, to large values that reach up to 100 m at the end of the
austral winter. Note that the PIRATA-SWE-derived MLD was not plotted at
19°S-34°W from 10 May to 11 November 2006 due to missing salinity
observations in the upper level. Perhaps the largest difference for the MLD
estimation between ROMS and in-situ data is noticed here at the 08°S-30°W
73
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
site (Figure 6.4a). Indeed, even if the seasonal evolutions are in phase, the
simulated MLD is generally systematically shallower (from 0 to about 30 m)
than the observational data. This feature is already visible in Figure 6.2,
where the ROMS 26°C isotherm depth at 08°S-30°W is always significantly
shallower. The systematic underestimation of the MLD at 08°S-30°W seems
to be related to the difficulty in estimating observational MLD. The vertical
measurements of salinity on the ATLAS moorings are presently limited to
only four levels (1, 20, 40, and 120 m) within the upper layer (see section 2.2
on Chapter 2). Another reason for this discrepancy can be related to the use
of climatological heat flux to forcing surface layer. One way to investigate
this last hypothesis is through a series of case studies to check the model’s
sensitivity to different forcings.
74
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
(a)
(b)
(c)
Figure 6.4. Comparison of temporal evolution of seasonal mixed layer depth (MLD)
at: (a) 8°S-30°W, (b) 14°S-32°W and (c) 19°S-34°W, provided by PIRATA-SWE in-situ
observations (dashed line) with ROMS (solid line) for the period of September 2005 to
June 2007.
75
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
6.3. Transport and pathways in the Southwestern tropical Atlantic
6.3.1. Meso-scale activity in the Southwestern tropical Atlantic
The averaged (2005-2007) simulated Sea Surface Height (SSH) and the
derived surface Eddy Kinetic Energy (EKE) of the study area are presented in
Figs. 6a-d. In these figures, the numerical results are compared to the
AVISO Rio05 product for the same periods. Mean SSH was obtained from
AVISO Rio05, combining hydrographic data, surface drifters velocities,
altimetry and a geoid model (Rio and Hernandez, 2004). Simulated and
measured surface Eddy Kinetic Energy (EKE) were calculated from SSH
gradients with a similar temporal sampling.
The ROMS-simulated boundary systems (Figures 6.5a and 6.5b,
respectively) bear good resemblance to those of AVISO Rio05 data (Figs. 6.5c
and 6.5d). The similarity of the isocontours as well as the eddy-induced
structures along the BC is particularly strong. The highest values are greater
than 50 cm
2
s
-2
, found in cSEC and NBUC/BC areas; this holds for both
AVISO Rio05 and ROMS. In general, observations show stronger meso-scale
activities in the Southwestern tropical Atlantic boundary as compared to the
ROMS simulation, but the geographic patterns are similar. It is possible to
identify three common characteristic areas of high surface variability that
are present in ROMS and AVISO Rio05 EKE charts. First, the near-shore BC
patch south of 16
o
S, where meso-scale cyclonic structures are documented
(Schmid et al., 1995; Campos, 2006). Second, along the Brazilian edge, north
of Salvador, following the NBUC jet, and finally the zonal band close to the
northernmost boundary of the integration domain (5-6
o
S), where the central
branch of the SEC develops (Lumpkin and Garzoli, 2005). Concerning the
coastal region, SSH (and consequently EKE) maps indicate that the
NBUC/BC proper is not well represented in AVISO Rio05 dataset. This is
probably associated with the partial inability of the altimetry data to
accurately capture part of the strong boundary current inertia near the
coastline.
76
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure 6.5. Averaged (2005-2007) SSH and EKE comparisons between ROMS
simulations and AVISO Rio05 data: (a) ROMS - SSH (cm), (b) ROMS - EKE (cm
2
s
-2
), (c)
RIO05+AVISO - SSH (cm) and (d) AVISO - EKE (cm
2
s
-2
).
77
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure 6.6. Transport function (Sv) for the modelled seasonal averaged (three
months, 2005-2007) currents integrated from 600 m to the sea surface for a) MJJ and
b) OND. The 1000 m isobath is represented by a red line.
78
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
6.3.2. Seasonal (2005-2007) variability of the sSEC/NBUC/BC system
The simulated volume transport in Sv (1 Sv = 10
6
m
3
s
-1
) of the
sSEC/NBUC/BC system between the surface and 600 m is shown in Figure
6.6a, b for the averaged period of MJJ and OND (2005-2007), respectively.
The first characteristic of ocean circulation expressed in Figure 6.6a, b is the
inclination of the transport isolines during MJJ compared to the OND
period. Indeed, during OND, when less intense SW trade winds are present,
the transport lines west of 24
o
W approach the continent at more or less
constant latitudes. In this case, the southern branch of the SEC, which
carries 5 Sv, reaches the coast at about 16
o
S, while the transport isoline of
10 Sv approaches the land near 10
o
S. Even during MJJ, characterized by
more intense action of the SW trades, the transport isolines of 5 and 10 Sv
are clearly tilted in relation to the parallels, approaching the edge of the
continental shelf at around 18
o
S and 14
o
S, respectively. Through these
figures, we can also see a strengthening of the alongshore NBUC transport
during the period in which the isolines reach the coast further south, with
simultaneous weakening of the transport of BC to the south (MJJ, Figure
6.7a). The inverse situation is observed in Figure 6.6b (OND), when the
bifurcation of sSEC occurs further north. In this case, the NBUC is
weakened and the BC is intensified (see also Figure 6.10a-b).
79
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure 6.7. Model simulation of meridional current and transport values averaged
from September 2005 to July 2007 along zonal sections (0-1500 m) at: (a) 8°S, (b)
14°S and (c) 19°S. Positive (negative) values indicated by solid (dashed) lines
correspond to northward (southward) currents.
80
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Meridional current and transport estimates averaged for the period of
September 2005 to July 2007 are presented here along zonal sections (0-
1500 m) at 8°S, 14°S and 19°S, which correspond to the latitudes of the
three PIRATA-SWE buoys (Figure 6.7a-c). Three
t
σ
levels are selected
according to the local dynamics. The first sigma level (
t
σ = 24.5 kg m
-3
)
separates the upper Tropical Surface Water (TSW) from the upper
thermocline waters. The second sigma level (
t
σ
= 26.8 kg m
-3
), which is
located at about 150 m, is the lower level of the water supplying the EUC
(Schott et al., 1998). The third one (
t
σ
= 32.15 kg m
-3
), which corresponds to
~1000 m, indicates the lower boundary of the upper warmer waters as well
as the lower boundary of the Antarctic Intermediate Water (AAIW). Below a
depth of 1100 m, we found the North Atlantic Deep Water (NADW) extending
to about 4000 m. This is the layer where the Deep Western Boundary
Current (DWBC) transports southward cold waters from the northern
hemisphere.
The zonal mean sections in Figure 6.7a-c show a clear representation of
the NBUC skirting the coastline from 100 to 1000 m, with a northward
transport increasing from 6.7 Sv at 19°S to 20.3 Sv at 8°S. Along the latitude
of 8°S, the core of the NBUC is located about 50 km from the coast, at
approximately 50-to-600 m depth (Fig. 8a), and between 250-to-650 m depth
at 14°S (Fig. 8b) and at 19°S (Figure 6.7c). Still, in the first 1500 m, the
presence of a mean southward flow east of the lower NBUC is observed from
the model transects at 8
o
S and 14
o
S (i.e. 150 and 350 km from the coast,
respectively), suggesting a continuous offshore recirculation branch at
depths between 200-1400 m. The southward transports of 2.3 Sv at 8°S and
4.3 Sv at 14°S (Figure 6.7a, b) were also found in the LADCP measurements
and EOF analysis performed by Schott et al. (2005) and Schuckmann (2006),
who found mean representative values of -5.2±4.9 and -4.1±3.7 Sv at 5°S
(ship sections) and 11°S (ship sections + mooring array), respectively. There
are three possible explanations for the origin of this flow. One could be the
deflection of zonal currents. However, no evidence of a southward deflection
has been reported in the literature for these currents, because the South
81
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Equatorial Undercurrent and the Southern Intermediate Countercurrent
transport stays practically constant between 35
o
W and 28
o
W (Schott et al.,
2005). A second explanation for this southward flow located east of the
NBUC is an inflow from the east as part of the deep cSEC, which deflects
southward before reaching the Brazilian coast, and thus forms an offshore
counterflow (hereinafter referred as CFLOW) to the deep NBUC.
Furthermore, the flow could be explained by a retroflection of the deep NBUC
just a little bit north of the northeastern tip of Brazil to supply the
southward offshore flow across 5
o
S, identified in the high resolution (1/12
o
)
MICOM model simulation (Schott et al., 2005). In order to investigate these
hypotheses, we plotted in Fig. 6.8 the annual averaged (2005-2007)
meridional current (cm s
-1
) and transport (in Sv) obtained by ROMS for: (a)
the along zonal section (0-1500 m) at 7°S, and (b) the depth range of 200-
1000 m, where southward CFLOW is stronger (see Fig. 6.7a and b). Despite
the proximity of the northern boundary to the integration domain (and as a
consequence, of the influence of the numerically-imposed sponge layer
condition between 5-6
o
S), our modeling results plead in favor of the second
case mentioned above, i.e., that the offshore counterflow to the deep NBUC is
fed by the southernmost limb of the deep cSEC.
Further east of the CFLOW (-2.3 Sv) shown in Figure 6.7a, we can
distinguish two main current systems, one to the north with a corresponding
transport of 2.1 Sv from 200 to 300 km away from the coast, and another to
the south (-3.5 Sv) located between 300 and 500 km from the shoreline.
These meridional offshore cores were also found by Schott et al. (2005)
between the 9
o
S and 11
o
S transects.
82
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
(a)
Figure. 6.8. Annual averaged (2005-2007) meridional current (cm s-1) and transport
(in Sv) obtained by ROMS for: (a) the along zonal section (0-1500 m) at 7°S. Positive
(negative) values indicated by solid (dashed) lines correspond to northward
(southward) currents, (b) the depth averaged range 200-1000 m, where CFLOW is
stronger
83
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
The southward-flowing BC, confined to the shallow and near-shore
part of the Brazilian continental slope, is especially recognizable at the 19°S
section with a mean transport of -2.8 Sv and meridional velocities ranging
from 10 to 20 cm s
-1
(Figure 6.7c). Miranda and Castro (1981) identified the
BC at 19°S as a surface narrow current (~ 75 km) limited to the upper 500
m. Evans et al. (1983) indicate that the BC remains confined and organized
as a coherent flow above the continental shelf at 20.5°S. The World Ocean
Circulation Experiment (WOCE) current-mooring measurements obtained at
19°S show a BC confined to the upper 200 m depth, with a mean southward
velocity of approximately 15 cm s
-1
(Müller et al., 1998). The mean cross-
shore section at 19°S obtained here from the 2005-2007 simulations (Figure
6.7c) confirms the presence of a southward BC tight flow (less than 100 km)
limited to the top 200 m. At 19°S east of the NBUC flow (6.7 Sv), one can find
the upper part of the DWBC transporting cold waters southward (1.8 Sv)
above
t
σ = 32.15 kg m
-3
(Fig. 6.7c).
In our simulation, the sSEC bifurcation reaches its southernmost
position in MJJ and its northernmost in OND, which corresponds
respectively to the months of July (southernmost) and November
(northernmost) found in the climatological runs of Rodrigues et al. (2007).
Furthermore, a time variation of the model outputs indicates (Figure 6.9)
that the NBUC system strengthens around May 2006 and May 2007, i.e.
when the sSEC bifurcation reaches its southernmost position, while the BC
transport is decreasing at that time until May 2006, when it is practically
null. On the other hand, maximum southward BC transports are verified
during January 2006 as well as in January 2007 and March 2007, just after
the period when sSEC bifurcation reaches its lowest latitudes, with a
minimum northward NBUC flow in December 2005 and October/December
2006 (Figure 6.9).
84
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure. 6.9. Interannual variability of the North Brazil Undercurrent (NBUC) transport
along 8oS (from the surface to 400 m), and Brazil Current (BC) transport along 19oS
(from the surface to 400 m), obtained from ROMS simulations for the period of
September 2005 to July 2007. Positive (negative) values indicated by the black (gray)
line correspond to northward (southward) transports.
The depth dependence of the sSEC bifurcation latitude is abridged in
Figure 6.10a, b. In these figures, the sSEC bifurcation is represented by a
white line, where the averaged (2005-2007) MJJ and OND meridional
velocity (spatially averaged within a 1° longitude band off the Brazilian coast)
is zero. The white areas in Figure 6.10a-b represent the Vitória-Trindade
Ridge and Abrolhos Bank. ROMS also agrees well with the numerical
findings of Harper (2000), Malanotte-Rizzoli et al. (2000) and Rodrigues et al.
(2007), indicating a poleward depth increase of the sSEC bifurcation along
the Brazilian coastline. Averaged MJJ results in Figure 6.10a show that the
bifurcation varies from near the surface at 13°S to 500 m depth at the model
boundary at 24°S. During this period, the NBUC is strengthened and the BC
weakened. In contrast, the OND averaged results indicate that the
bifurcation shifts southward from 8°S at the surface layers to 20°S at 500 m
depth, when a weaker NBUC and stronger BC are observed (Figure 6.10b).
This southward deepening of sSEC bifurcation along the shoreline results in
85
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
a similar poleward depth increase of the NBUC cores, as verified in the 8°S,
14°S and 19°S transects (Figures 6.7a-c). These numerical results have also
been detected in field measurements. Stramma and England (1999) and
Rodrigues et al. (2007) showed from hydrographic data that the SEC
bifurcation takes place at 14-16°S in a near-surface layer (top 100 m), then
at 14-20°S in the layer between 100-500 m, and at 21-26°S in the AAIW
(500-1200 m). Wienders et al. (2000) also used hydrographic data to indicate
that the SEC bifurcation latitude is 14°S at the surface, 24°S in the 400-500
m layer and around 26°S-28°S in the AAIW. More recently, the annual mean
dynamic height and geostrophic flow charts generated by Rodrigues et al.
(2007) from CTD and bottle data (Curry, 1996) confirmed the poleward shift
of the sSEC with increasing depth.
86
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure. 6.10. Seasonal averaged (three months, 2005-2007) meridional velocity (m s-
1) obtained from the ROMS simulation for a) MJJ and b) OND. The velocities are
averaged within a 1° longitude band off of the Brazilian coast. The white line is the
contour of zero velocity that represents the bifurcation of the sSEC. The white areas
represent the Vitoria-Trindade Ridge and Abrolhos Bank.
87
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
6.3.3 Coupling westward sSEC and wind stress
The zonal current and transport averaged from September 2005 to
July 2007 along the PIRATA-SWE array indicate how the southern part of
the SEC extends before reaching the western continental boundary (Figure
6.11a). Indeed, we note a complex succession of more powerful westward
systems. From south to north, we find two near-surface cores of -8.5 and -
2.9 Sv, corresponding to the broad and relatively weak sSEC westward flow
between 10°S and 19.5°S (Stramma, 1991; Stramma and Schott, 1999;
Lumpkin and Garzoli, 2005). More to the north, in the vicinity of the
northern PIRATA mooring, a third westward core of 2.8 Sv surfaces as the
southernmost part of the cSEC (Lumpkin and Garzoli, 2005). ROMS results
also point out a less powerful near-surface eastward core of 1.25 Sv located
between sSEC and cSEC. Other subsurface narrow eastward circulations
may reach the whole water column between 200 and 1200 m depth (not fully
shown in Fig. 6.11a).
88
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
Figure 6.11. (a) Mean annual volume transport averaged across the section along the
PIRATA-SWE array between 8°S-30°W and 19°S-34°W. Positive (negative) values indicated by
the solid (dashed) lines correspond to eastward (westward) currents. The two horizontal solid
black lines indicate the 24.5 and 26.8 sigma-t values (in kg m-3), respectively. (b) Time
evolution (Sep. 2005-June 2007) of the westward SEC transport (0-400 m depth) obtained
from ROMS simulations along the PIRATA-SWE array (solid line), and the QuickSCAT zonal
wind speed anomalies over the model domain (dashed line). The variance spectra for: (c) the
zonal wind, and (d) westward sSEC transport. The confidence levels are indicated by shaded
areas in both spectra.
89
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
According to this ROMS simulation, the mean westward transport
between 8°S-30°W and 19°S-34°W, corresponding to the sum of the sSEC
and cSEC, is -14.9 Sv for the upper 400 m, which agrees well with previous
observations and numerical studies. For instance, the hydrographic data
used by Stramma (1991) showed a westward SEC transport of about -20 Sv
along 30°W for the upper 500 m comprised between 3°S and 19°S.
Furthermore, the recent climatological simulations of Rodrigues et al. (2007)
indicate an annual mean westward SEC transport between 6°S and 22°S of -
15 Sv along the 30°W meridian and down to 400 m. These authors found
maximum SEC transports during JFM of about -15.6 Sv, and minimum
flows of -14.0 Sv during MJJ. From September 2005 to July 2007, monthly-
averaged westward transport (0-400 m depth) obtained from the ROMS
simulation across the PIRATA-SWE moorings (Fig. 6.11b), shows large
interannual and intraseasonal variability. The weakest westward transport (~
10-11 Sv) is noted in mid-2006, while the strongest westward SEC transport
(~ 18-19 Sv) occurs in February 2006, and three times from September 2006
to June 2007. The time evolution (Sep. 2005-Jun. 2007) of the QuickSCAT
zonal wind speed anomalies over the model domain is also plotted in Figure
6.11b. Since the mean zonal wind for this region is negative, a positive
difference indicates weaker westward zonal winds than the mean, while
negative deviation values indicate monthly averaged winds greater than the
mean for the simulation period. Positive zonal wind anomalies are associated
with a decrease in the upper transport of SEC, with time lags between zero
and two months. These values occur in April 2006, July 2006, November
2006, February 2007 and April 2007 (Figure 6.11b). Minimal SEC transports
are observed in May 2006, August 2006, January 2007 and April 2007. On
the other hand, the negative zonal wind anomalies in Figure 6.11b
(December 2006, January 2006, August 2006, September 2006, January
2007 and March 2007), are followed by an increase in the SEC transport,
which is observed in February 2006, September 2006, February 2007 and
April 2007. In Figure 6.11c-d the zonal wind speed anomalies over the model
domain are correlated with westward sSEC transport (0-400 m depth)
obtained from the ROMS simulation across the PIRATA-SWE (Sep. 2005 to
90
Chapter 6 – Interannual high-resolution regional ocean dynamics simulation …
91
Jun. 2007). Since the interest here is in the seasonal signals, a 30-180 day
band-pass filter was applied to both the atmospheric and the modeling data
to be compared. For the cross-correlation field, the variance-preserving
spectra are constructed for the area with higher cross-correlation. The
confidence levels are indicated by shaded areas in both spectra. The spectral
analysis points out the signal periodicity of both datasets. The calculated
maximum cross-correlation between zonal wind and westward transport is
0.51. The spectral analyses show two dominant signals at approximately 90
and 50 day periods for both zonal wind speed and sSEC transport (Figure
6.11c and d, respectively). In addition to seasonal effects, these ocean
signals may also represent the westward propagation of Rossby waves, as
evidenced in the mooring measurements and analyses of Schott et al. (2005)
and Schuckmann (2006).
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
Chapter 7
7. Regional biogeochemical modelling
in the Southwestern tropical Atlantic
In the oligotrophic regions of the ocean, the supply of inorganic nutrients in
the euphotic layer may limit the concentration of microalgal biomass, the
rate of phytoplankton growth, or both (Marañón et al., 2000). Several
mechanisms have been proposed to explain the formation and maintenance
of the deep chlorophyll maximum (DCM), which is an oceanographic feature
of tropical and subtropical oceans. Higher in-situ growth at the nutricline
rather than in the upper mixed layer, and acclimation of phytoplankton to
low irradiance, has been suggested (Cullen,1982) as the cause of DCM
formation on subtropical gyres.
Early reports of low productivity in the subtropical gyres (Steeman-
Nielsen and Jansen, 1957; Thomas, 1970; Eppley et al., 1973) have been
considered to be in error due to methodological deficiencies (Fitzwater et al.,
1982). Nevertheless, recent work carried out in oligotrophic regions, using
new techniques (Malone et al. 1993; Letelier et al. 1996), presents a large
range in the rates of phytoplankton production and growth inside these
areas. In fact, this controversy will be maintained, while the lack of
92
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
knowledge on temporal and spatial variability in the biology of open ocean
still exists (Marañón et al., 2000). Consequently, any uncertainties in the
productivity estimates in these areas will have significant effects on the
prediction of global biogeochemical models (Marañón et al., 2000). On the
other hand, there are few studies on the physical processes related with
DCM formation inside the Southwestern tropical Atlantic. Montes (2003)
suggests that the support of nutrient inside the euphotic zone, in the
Southwestern tropical Atlantic, is maintained by diffusion process in
nutricline basis. In this chapter, we use a coupled physical-biogeochemical
modelling approach to increase the knowledge of physical processes that
contribute to the formation of DCM, and investigate their seasonal variability
related with SEC/NBUC/BC system current.
7.1 The coupled ROMS approach
The ROMS version used in this chapter was previously presented in
Chapter 3. This modelling was based on physical approach of climatological
forcings at the Southern tropical Atlantic grid (Figure 7.1) adopted in
Chapter 5. The initial condition to start the biogeochemical model was
obtained from WOA2005 (Conkright et al., 2002) for nutrient and from
seasonal and annual SeaWifs data (Penven et al., 2008; O’ Reilly et al., 1998)
for Chlorophyll a, phytoplankton and zooplankton concentrations inside the
euphotic layer.
93
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
Figure 7.1. Model domain (dashed gray lines) with the PIRATA-SWE locations (filled
triangles). Section along three PIRATA-SWE sites (solid lines). Thin dashed lines are
100m and 1000m isobaths, respectively.
7.2 Evaluation of simulated and satellite Chlorophill a data
As a preliminary test to validate the numerical results, the model
accuracy was evaluated by comparing monthly mean of chlorophyll a surface
maps derived from the ROMS model to monthly averaged 2001-2006 daily
SeaWifs Chlorophyll a data. This comparison was done for two different
scenarios: November, which corresponds to the austral spring (Figure
7.2a,c), and July, which corresponds to the austral winter (Figure 7.2b,d).
In the region of subtropical gyre, the model and SeaWifs results are
consistent, showing a low concentration of chlorophyll a in these two months
that correspond to oligotrophic areas (Figures. 7.2a, c and 7.2b,d). However
the bloom caused by upwelling on the south Brazilian coast is not in phase
with SeaWifs maps. The higher concentrations along the coast occur in July
to SeaWifs (Figure 7.2b) as the ROMS result indicates in November. It
suggest that constant primary production and growth rates does not
94
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
represent locally the near coast upwelling. Furthermore, along the north
boundary (5ºS) ROMS indicates concentrations of chlorophyll a that does not
agree with satellite images. These vortices with blooms discrepancies
indicates that boundary conditions applied on NCPZD model can take more
influence inside domain that the thickness of relaxation layer of open
boundaries, mainly at north boundary.
(a)
(b)
(c)
(d)
Figure 7.2. Chlorophyll a distribution on surface to SeaWifs on: a) November b) July
and ROMS results on: c) November d) July
95
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
7.3 Salinity intrusion causes and implications
Figure 7.3a presents monthly averaged results of simulated salinity to
Maximum Salinity Waters (MSW, hereinafter) values of salinity (salt
concentration > 37.5) occur south of 14ºS, a region inside the subtropical
gyre, with high surface salty waters. At mid-latitude the negative fresh water
balance is the main process that contributes to the increase in salinity by
high evaporation rates and low rain quantity. Furthermore, the surface
forced wind that transports surface waters masses to north are weaker in
November than in July. This holds the spreading towards the north of these
waters. Figure 7.3b shows an intrusion of MSW between 50 and110 meters
until it reaches near 10ºS. The equatorward transport of MSW is due to wind
driven circulation (Rodrigues et al., 2007). It transfers surface MSW of
subtropical gyre to lower latitudes. This presented result also suggests that
the subduction of MSW (finger) and northward spreading, in July, is
associated with the presence of lower density tropical water ( σ
t
< 24.8) at
low-latitudes in the SAWP region. The seasonal variability of MSW along this
section causes the occurrence of isohaline layer shallower (deeper) than the
isothermal layer in November (July). This has been studied, in the tropical
ocean, since the Meteor research cruises in 1936 (Defant, 1961). The barrier
layer (BL), which is located between the halocline and the thermocline
(Lukas and Lindström, 1991), may isolate the upper isohaline layer from the
cold and rich nutrient waters.
96
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
(b)
(a)
Figure 7.3. Vertical salinity distribution of ROMS along PIRATA-SWE sites on: a)
November b) July.
97
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
7.4 Seasonal Nitrate concentration
The monthly averaged nitrate results of the biogeochemical model in
November and July (Figures 7.4a,b respectively) shows that there doesn’t
exist great differences between nitrate distributions along this section.
Although, small differences that appears in the inclination of nutricline
between 14ºS and 8ºS, following the 36.5 isoline of salinity, shows best
agreement with Marañón et al. (2000) results of south subtropical gyre. This
small nutricline displacement to a deeper level between the two northern
PIRATA-SWE sites, in July, showing more nutricline establishment, may be
caused by intrusion of surface and poor nutrient MSW, which comes from
mid-latitudes and subductes until it reaches near 10ºS (Figure 7.3b). The
presence of a null concentration line along the section around σ
t
=26.8, July
(Figure 7.4b), indicates that the subduction of poor nutrient MSW, may
decrease the nutrient support on upper layers. Besides, in July (austral
winter) intensified trade winds, in this region, causes enlargement MLD. On
the other hand, in November (Figure 7.4a) at late austral spring, a small and
shallow thermocline appears between 19ºS and 14ºS, due to more intensities
of heat flux and weaker trade winds, the null line of nitrate concentration
almost disappears. However, low concentrations still are achieved along the
MLD. This agrees with Montes (2003) results, suggesting that the major
source of nutrients comes from deeper rich nutrient water by diffusion
processes.
98
Chapter 7 – Regional biogeochemical modelling in Southwestern tropical …
99
Figure 7.4. Vertical nitrate distribution of ROMS along PIRATA-SWE sites on: a)
November b) July.
Chapter 8 - Summary and perspectives
Chapter 8
8. Conclusions and perspectives
The tropical South Atlantic is a region of complex links between climatic
variability of the sea surface temperature (SST) and its heat content in upper
layers to an atmospheric convective system as precipitation on the adjacent
continent. The main goal of this thesis was to investigate the seasonal and
intraseasonal variability of mass and heat transport in two different regions
of the South Tropical Atlantic: the equatorial band and the Southwestern
tropical Atlantic. In order to achieve the proposed goal, a new generation
state-of-the-art ROMS was applied to reproduce the dynamic of these
100
Chapter 8 - Summary and perspectives
systems.
Numerical transports and currents results were compared to ocean
measurements obtained from oceanography cruises. These comparisons
suggest that the climatological modeling approach employed here reproduces
the main features of the zonal circulation observed from field measurements.
For example, numerical currents and transports distributions showed, in
Chapter 4, a very neat distinction between eastward and westward flows
along the equator. The model stressed the presence of a EUC core with
velocities reaching 0.8 m.s
-1
, occupying a deeper position in August (80-110
m depth) and shallower in March (50-80 m depth). The transport of EUC
ranged from 7.9 Sv (March, at the 15°W cross-equatorial section) to 15 Sv
(August, at 25°W). These numerical findings are confirmed by the sea
measurements and analysis of Stramma et al (2003), Giarolla et al. (2005)
and Brandt et al. (2006). The westward flows of the SECc and the SECn were
also well represented by this approach. The transport of SECc ranged from -
2.8 Sv (August, at 15°W) to -12 Sv (March, at 25°W), while the SECn
intensity ranged from -1.9 Sv (August, at 25°W) to -11.9 Sv (March, at 15°W).
Closer to the African shore, the seasonal variation of the equatorial currents
transport along the 5°W section shows highest values of the EUC and GC in
the period April–August, and lowest values occurring in the period
September-March. On the other hand, the SEC seasonality presents inverse
standards, i.e., weaker SECc and SECn transports in April–August and
stronger westward flows during the period September–March. The thermal
structure of the upper 500 m obtained in simulations was compared with
PIRATA Program dataset measured at the equatorial moorings. Special
101
Chapter 8 - Summary and perspectives
interest can be seen in the very well reproduced MLD for the equatorial
PIRATA sites. In particular, the upwelling induced shallower MLD verified by
the eastern buoys during austral winter and late spring. The Hovmüller
diagrams indicated that the vertical distribution of the simulated
temperature agrees with field measurements. Best agreement between model
and sea measurements is achieved close the surface on the western part of
the Atlantic basin (PIRATA sites 35°W and 23°W). Larger differences (~ 1 to
3°C) are noted between 100 and 300 m for eastern PIRATA sites (10°W and
0°W) in austral winter. The reduced flow in zonal currents observed (and
numerically reproduced) in August is associated to the intensification of the
equatorial upwelling mechanism in the eastern basin (Stramma and Schott,
1999). These authors also remind us that an equatorial upwelling and
downwelling exist on the equatorial currents system but where it takes place
exactly is an unsolved question. It seems to be a next interesting research
step to be explored with the numerical simulation approach. The seasonal
evolution of the Southwestern Atlantic Warm Pool (SAWP) – Cold Tongue
system was accurately represented on simulation. This is a very important
modelling issue since the eastern part of the Brazilian Northeastern Region
is subject to the effects of equatorial easterly atmospheric waves emanating
from convection over Africa. The generalized flooding conditions over the
coast of this region are observed when pulses of easterly waves encounter
warmer SST in the SAWP area (Nobre et al., 2004). These results are very
encouraging. The ROMS seems to be capable of reproducing the main
features of the mean EUC/NECC/GC and SECc/SECn at different sections
along the equator. The verified model adjustment to offshore field data is
102
Chapter 8 - Summary and perspectives
promising. For example, these results can be extensively used in the future
as boundary conditions for smaller-scale sea modelling, for evaluating the
remote influences of the open ocean forcing over the dynamics of the still
little studied oceanic islands in the northeastern Brazilian waters.
In Chapter 5, a climatological modelling brings new insights to Oceanic
terms (advection + diffusion) responsible for the seasonal evolution of the
heat budget inside the upper ocean mixed layer (MLD), which was calculated
from numerical outputs and compared with the atmospheric forcing balance.
Local illustrations of these results are given for the PIRATA-SWE sites and
along the western Atlantic Ocean boundary. It is confirmed that the
atmospheric surface fluxes are the main forcing mechanism acting on the
heat budget in upper ocean mixing layers. Nevertheless the oceanic
contribution is also very important, and can compensate the atmospheric
forcing effect in the southern part of the study region. During practically all
year, the oceanic component acts as an inversed contribution of the
atmospheric forcing. During the warm season, that corresponds for instance
(i.e. positive effect of the atmosphere) to a relatively strong cooling of the
mixing layer mainly by vertical diffusion through the thermocline from colder
deeper waters, and, during the cold season (i.e. negative effect of the
atmosphere) to a weaker warming of the surface waters, mainly by horizontal
advection.
These ROMS outputs, in this region, confirm the extreme complexity of
the oceanic circulation in this area. Along the virtual track represented by
the three PIRATA moorings (about 1400 km), we highlighted no less than six
westward currents and four eastward currents on a yearly average. Although
103
Chapter 8 - Summary and perspectives
the annual net mass transport across this PIRATA track is normally directed
towards the west (18.8 Sv for the upper 600 m, in agreement with early
estimations), that indicates that it will not be possible to perform any
geostrophic computations using only the density profiles measured by the
PIRATA sites. This implies that a continuous observation of the ocean
dynamics in this area will be necessary to implement the PIRATA network
with other better adapted systems.
Chapter 6 presents the use of ROMS forced by interannual boundary
conditions in order to investigate the intraseasonal and interannual
variability in the southwestern Atlantic for the 2005-2007 period. After
checking that ROMS makes it possible to capture correctly the meso-scale
phenomena illustrated by instant SST patterns, numerical temperature
values, issued from ROMS simulation, were compared to vertical profiles
obtained by the first two years of available data from the recent deployed
PIRATA-SWE buoys. Comparison between simulated temperature and in-situ
temperature data are in good agreement for the first 500 m ocean layer.
These ROMS outputs confirm the extreme complexity of the oceanic
circulation in this area. This is obviously the case along the western
continental boundary where many alongshore currents coexist and can
interact. The interannual numerical results indicate a close relationship
between SEC and the western boundary currents flowing along the Brazilian
edge. When the sSEC bifurcation reaches its southernmost position (MJJ),
the northward NBUC transport is stronger (May 2006, May 2007) and the
BC transport decreases. Otherwise maximum southward BC flows are
verified during January 2006 and January/March 2007, with minimum
104
Chapter 8 - Summary and perspectives
northward NBUC flows in December 2005 and October/December 2006,
during the period when sSEC bifurcation reaches its lowest latitudes (OND).
Along the section delineated by the PIRATA-SWE moorings, we
highlighted the existence of three strong near surface westward currents
alternated by two weak eastward currents on a yearly average. Although the
annual net mass transport across this PIRATA track is usually westward
(14.9 Sv for the upper 400 m), this indicates that geostrophic computations
using only the density profiles measured by the three PIRATA-SWE sites are
somewhat misfit. This situation is worsened because the vertical
measurements of salinity on the ATLAS moorings are presently limited to
only four levels within the 0-120 m upper layer. This implies that for a
continuously observation of the ocean dynamics in this area it will be
necessary, not only to upgrade the vertical sampling of the PIRATA-SWE
network, but, to also certainly implement other adapted systems.
The evaluation of this high resolution regional ocean simulation shows
the capability of the model in reproducing the known ocean dynamics in the
region and their variability. One can note that it has been achieved on a
relatively long period and as a free run (i.e. without the need of artificially
restoring the model solution towards observations). Hence, we can analyze
now this model solution to rigorously diagnose the dynamical balances and
the eddy dynamics.
Chapter 7 presents results of an ecological model that runs coupled
with the Climatological ROMS, as set in Chapter 5. Evaluation of monthly
mean chlorophyll a between simulated and SeaWifs satellite data on two
105
Chapter 8 - Summary and perspectives
different months shows that the model can reproduce main features of
primary production in an oligotrophic area limited by nutrients support,
where small oscillation in this support can produce high difference in
results. The main contribution of this modelling to knowledge of
biogeochemical processes around Southwestern tropical Atlantic occurs
when this simulation is capable of representing a seasonal wind driven
transport. That shows increases in subduction of MSW in July. This comes
poor nutrient shallow waters to the thermocline basis. On other hand, small
differences between nutrient distribution along nutricline in November and
the maintenance of low primary production inside the euphotic zone along
the year at the subtropical gyres, confirmed by Marañón et al. (2000),
suggest that the vertical diffusion of nutrients acts to balance a deeper
mixed layer and decreasing in light availability in July. The use of ROMS
coupled with NCPZD could be used in the future as a useful tool to estimate
the dispersion of the tuna eggs and larvae in the divergence of SEC (Mullon
et al., 2002; Hugget et al., 2003).
A potential outcome of this thesis is the possibility of coupling the
ROMS with a dynamically active regional atmospheric model, when SST,
heat and freshwater fluxes are calculated from processes which couple both
marine boundary layers. These studies are underway. The ROMS can be
used with a higher resolution to resolve smaller scales, using improved
turbulence closure schemes, or using better surface forcing or lateral
boundary conditions. Using the nesting capability of ROMS (Penven et al.,
2006), we could also explicitly resolve fast propagating signals across the
whole Tropical Atlantic while addressing the coastal response in the
106
Chapter 8 - Summary and perspectives
107
Southwestern region.
References
Allen, J. S., 1980. Models of wind-driven currents on the continental shelf.
Annual Review of Fluid Mechanics 12, 389 - 433.
Arhan, M., Mercier, H., Bourlès, B., Gouriou, Y., 1998. Hydrographic
sections across the Atlantic at 7°30N and 4°30S. Deep Sea Research I
45, 829-872.
Battisti, D. S., Hickey, B. M., 1984. Application of remote wind-forced
coastal trapped wave theory to the Oregon and Washington coasts.
Journal of Physical Oceanography 14, 887 - 903.
Blaas, M., Dong, C. M., Marchesiello, P., McWilliams, J. C., Stolzenbach, K.
D., 2007. Sediment-transport modeling on Southern Californian shelves:
A ROMS case study. Continental Shelf Research 27, 832-853.
Blanck, H. F., 1999. Using TOPEX Satellite El-Niño altimetry data to
introduce thermal expansion and heat capacity concepts in chemistry
courses. Journal of Chemical Education 76, 1635-1646.
Blanke B., Arhan M., Madec G., Roche, S., 1999. Warm water paths in the
equatorial Atlantic as diagnosed with a general circulation model.
Journal of Physical Oceanography 29, 2753-2768.
Bourlès, B., Molinari, R. L., Johns, E., Wilson, W. D., Leaman, K. D., 1999a.
Upper layer currents in the western tropical North Atlantic (1989–1991).
Journal of Geophysical. Research 104(C1), 1361–1375.
Bourlès, B., Gouriou, Y., Chuchla, R., 1999b. On the circulation in the
upper layer of the western equatorial Atlantic. Journal of Geophysical.
Research 104, 21151–21170.
Bourlès B., Lumpkin, R., McPhaden, M. J., Hernandez, F., Nobre, P.,
Campos, E., Yu, L., Planton, S., Busalacchi, A. J.,Moura, A. D., Servain,
J., Trotte, J., 2008. The PIRATA Program: History, Accomplishments,
and Future Directions. Bulletin of the American Meteorological Society 89,
1111-1125. doi: 10.1175/2008BAMS2462.1
Brainerd, K. E., Gregg, M. C., 1995. Surface mixed and mixed layer depths.
108
Deep-Sea Research I 42, 1521-1543.
Brandt P, Schott FA, Provost C, Kartavtseff A, Hormann V, Bourlès B and
Fischer J. 2006. Circulation in the central equatorial Atlantic: Mean and
intraseasonal to seasonal variability, Geophysical Research Letters 33,
L07609, doi: 10.1029/2005GL025498.
Brink, K. H., 1982. A comparison of long coastal trapped wave theory with
observations off Peru. Journal of Physical Oceanography 12, 897 – 913.
Brink, K. H., Stuart, D. W., Van Leer, J. C., 1984. Observations of the
coastal upwelling region near 348309N off California: Spring 1981.
Journal of Physical Oceanography 14, 378 – 391.
Brink, K. H., 1989. Evidence for wind-driven current fluctuations in the
western North Atlantic. Journal Geophysical Research 94, 2029 - 2044.
Brink, K. H., 1991. Coastal-trapped waves and wind-driven currents over the
continental shelf. Annual Review of Fluid Mechanics 23, 389 - 412.
Camayo, R., Campos, E. J. D., 2006. Application of wavelet transform in the
study of coastal trapped waves off the west coast of South America.
Geophysical Research Letters 33, L22601, 1-5.
Campos, E. J. D., Goncalves, J. E., Ikeda, Y., 1995. Water mass
characteristics and geostrophic circulation in the south Brazil Bight -
summer of 1991. Journal of Geophysical Research 100(9), 18537 -
18550.
Campos, E. J. D., Piola, A. R., Miller, J. L., 1999. Water mass distribution on
the shelf and shelf-break upwelling in the southeast Brazil Bight. Annals
of the 10th Symposium on Global Change Studies – American
Meteorological Society, 9 - 12.
Campos, E. J. D., Velhote, D., da Silveira, I. C. A., 2000. Shelf break
upwelling driven by Brazil Current cyclonic meanders. Geophysical
Research Letters 27, 751 - 754.
Campos, E. J. D., 2006. The equatorward translation of the Vitoria eddy in a
numerical simulation. Geophysical Research Letters 33, L22607,
doi:10.1029/2006GL026997.
Carbonel, C. A. A. H., 2003. Modelling of upwelling–downwelling cycles
caused by variable wind in a very sensitive coastal system. Continental
109
Shelf Research 23,1559-1578.
Carton, J. A., Zhou, Z.,1997. Annual cycle of sea surface temperature in the
tropical Atlantic Ocean. Journal of Geophysical Research 102, 27813-
27824.
Castelão, R. M., Campos, E. J. D., Miller, J. L., 2004. A modelling study of
coastal upwelling driven by wind and meanders of the Brazil Current.
Journal of Coastal Research 20, 662-671.
Castelão, R. M., Barth, J. A., 2006. Upwelling around Cabo Frio, Brazil: The
importance of wind stress curl. Geophysical Research Letters 33,
L03602, 1 - 4.
Castro Filho, B. M., 1985. Subtidal response to wind forcing in the South
Brazil Bight during winter. Ph.D. Thesis. University of Miami, USA,
211pp.
Castro, B. M., Miranda, L. B., 1998. Physical oceanography of the western
Atlantic continental shelf located between 4
o
N and 34
o
S. In: Robinson,
A. R. and Brink, K. H. The Sea 11. John Wiley, New York, 209-251.
Cavalcanti, I. F. A., Kousky, V.E., 2003. Climatology of South American cold
fronts. In: VII International Conference on Southern Hemisphere
Meteorology and Oceanography, Wellington, New Zealand, 2003.
Chaves, R., R., Nobre, P., 2004. Interactions between sea surface
temperature over the South Atlantic Ocean and the South Atlantic
Convergence Zone. Geophysical Research Letters 31, L03204.
doi:10.1029/2003GL018647.
Chelton, D. B., Schlax, M. G., Witter, D. L., Richman, J. G., 1990. Geosat
altimeter observations of the surface circulation of the Southern Ocean.
Journal of Geophysical Research 95(C10), 17877 - 17903.
Church, J. A., Freeland, H. J., 1987. The energy source for the coastal
trapped waves in the Australian coastal experiment region. Journal of
Physical Oceanography 17, 289-300.
Cirano, M., Campos, E. J. D., 1996. Numerical diagnostic of the circulation
in the Santos Bight with COROAS hydrographic data. Brazilian Journal
of Oceanography 44, 105-121.
Cirano, M., Mata, M. M, Campos, E. J. D, Deiró, N. F. R., 2006. A circulação
110
oceânica de larga-escala na região oeste do Atlântico Sul com base no
modelo de circulação global OCCAM. Brazilian Journal of Geophysics
24(2), 209 – 230.
Conkright, M. E., Locarnini, R. A., Garcia, H. E., O'Brien T. D., Boyer, T. P.,
C. Stephens, C., Antonov, J. I. ,2002. World Ocean Atlas 2001: Objective
Analyses, Data Statistics, and Figures, CD-ROM Documentation,
National Oceanographic Data Center, Silver Spring, MD.
Cullen, J. J., 1982. The deep chlorophyll maximum: comparing vertical
profiles of chlorophyll a. Canadian Journal of Fisheries and Aquatic
Sciences 39 (5), 791-803.
daSilva, A. M., Young, C. C., Levitus, S., 1994. Atlas of surface marine data
1994, Vol. 1, algotithms and procedures, NOAA Atlas NESDIS 6. U.S.
Department of Commerce, NOAA, NESDIS, USA, 74 p.
Davey, M., Huddleston, M., Sperber, K., Braconnot, P., Bryan, F. &
coauthors, 2001. STOIC, a study of coupled model climatology and
variability in tropical ocean regions. Climate Dynamics 18, 403-420.
Defant, A., 1941. Die absolute Topographie des physikalischen
Meeresniveaus un der Druckflächen, sowie die Wasserbewegungen im
Raum des Atlantischen Ozean. Wissenschaftliche Ergebnisse Deutschen
Atlantischen Expedition auf dem Forschungs und Vermessungsschiff.
“Meteor” 1925-1927, 6(2), 191-260.
Dengler, M., Schott, F. A., Eden C., Brandt, P., Fischer, J., Zantopp, R.,
2004. Break-up of the Atlantic deep western boundary current into
eddies at 8°S. Nature 432, 1018–1020.
Enfield, D. B. and Allen, J. S., 1980. On the structure and dynamics of
monthly mean sea level anomalies along the Pacific Coast of North and
South America. Journal of Physical Oceanography 10, 557-578.
Enfield, D. B., 1987. The Intraseasonal oscillation in eastern Pacific sea
levels: How is it forced? Journal of Physical Oceanography 17, 1860 -
1876.
Eppley, R. W., Renger, E. H., Venrick, E. L., Mullin, M. M., 1973. A study of
plankton dynamics and nutrient cycling in the central gyre of the North
Pacific Ocean. Limnology and Oceanography 18, 534-551.
111
Evans, D. L., Signorini, S. R., Miranda, L. B., 1983. A note on the transport
of the Brazil Current. Journal of Physical Oceanography 13, 1732-1738.
Fennel, W., Lass, H. U., 2007. On the impact of wind curls on coastal
currents. Journal of Marine Systems 68, 128-142.
Fischer, J., Schott, F. A., 1997. Seasonal transport variability of the deep
western boundary current in the Atlantic. Journal of Geophysical
Research 102, 27751-27769.
Fitzwater, S. E., Knauer, G. A., Martin, J. H., 1982. Metal contamination and
its e!ect on primary production measurements. Limnology and
Oceanography 27, 544-551.
Foltz G., R., Grodsky, S., A., Carton, J., A., McPhaden, M., J., 2003.
Seasonal mixed layer heat budget of the tropical Atlantic Ocean. Journal
of Geophysical Research 108, 3146. doi:10.1029/2002JC001584.
Ganachaud, A., 2003. Large-scale mass transport, water mass formation,
and diffusivities estimated from World Ocean Circulation Experiment
(WOCE) hydrographic data. Journal of Geophysical Research 108, L3213,
doi:10.1029/2002JC001565.
Garzoli, S. L., Garrafo, Z., 1989. Transports, frontal motions and eddies at
the Brazil-Malvinas currents confluence. Deep Sea Research 36, 681-
703.
Giarolla E, Nobre P, Malagutti M and Pezzi LP. 2005. The Atlantic Equatorial
Undercurrent: PIRATA observations and simulations with GFDL Modular
Ocean Model at CPTEC. Geophysical Research Letters 32, L10617.
doi:10.1029/2004GL022206.
GODAE High Resolution Sea Surface Temperature Pilot Project - GHRSST-
PP, 2007. http://www.ghrsst-pp.org/, on line dataset, 04 June 2007.
Goes, M., Molinari, R., Silveira, I., Wainer, I., 2005. Retroflections of the
North Brazil Current during February 2002. Deep Sea Research I 52,
647-667.
Goni, G. J., Baringer, M. O., 2002. Surface currents in the tropical Atlantic
across high density XBT line AX08. Geophysical Research Letters 29,
2218. Doi:10.1029/2002GL015873.
Gordon, A. L., 1986. Interocean exchange of thermocline water. Journal of
112
Geophysical Research 91, 5013-5046.
Gordon, A. L., Greengrove, C. L., 1986. Geostrophic circulation of the Brazil-
Falklands confluence. Deep Sea Research I 33, 573-585.
Gould, R. W., 1987. The deep chlorophyll maximum in the world ocean: a
review. The Biologist 66(1-4), 4-13.
Grodsky, S. A., Carton, J. A., 2003. The intertropical convergence zone in the
South Atlantic and the equatorial cold tongue. Journal of Climate 16,
723-733.
Haidvogel, D. B., Arango, H. G., Hedström, K., Beckmann, A., Malanotte-
Rizzoli, P., Shchepetkin, A. F., 2000. Model evaluation experiments in
the North Atlantic basin: simulations in nonlinear terrain-following
coordinates. Dynamics Atmosphere and Oceans 32, 239-281.
Harper, S., 2000. Thermocline ventilation and pathways of tropical-
subtropical water mass exchange. Tellus 52A, 330-345.
Harrison, D. E., 1989. On climatological monthly mean wind stress and wind
stress curl fields over the world ocean. Journal of Climate 2, 57-70.
Hayes, S. P., L. J. Mangum, J. Picaut, A. Sumi, Takeuchi, K., 1991. TOGA-
TAO: A moored array for real-time measurements in the tropical Pacific
Ocean. Bulletin of the American Meteorological Society 72, 339-347.
Huang, B., Carton, J. A., Shukla, J., 1995. A numerical simulation of the
variability in the tropical Atlantic Ocean, 1980-88. Journal of Physical
Oceanography 25, 835–854.
Hugget, J. P., Freón, P., Mullon, C., Penven, P.,2003. Modelling the transport
success of anchovy Engraulis encrasicolus eggs and larvae in the
southern Benguela: The effect of spatio-temporal spawning patterns.
Marine Ecology Progress Series 250, 247-262.
Ikeda, Y., STEVENSON, M.,1978. Time series analysis of NOAA-4 sea surface
temperature (SST) data TOGA-TAO: A moored array for real-time
measurements in the tropical Pacific Ocean. Remote Sensing of
Environment 7, 349-362.
Jochum, M., Malanotte-Rizzoli, P., Busalacchi, A., 2004. Tropical instability
waves in the Atlantic ocean. Ocean Modelling 7, 145-163.
Jolliffe, I. T, 2002. Principal component analysis. Springer. New York.
113
Kayano, M. T., Rao, V. B., Andreoli, R. V., 2005. A review of short-term
climate variability mechanisms. Advances in Space Research 35, 843–
851.
Koné, V., Machu, E., Penven, P., Andersen, V., Garçon, V., Fréon, P.,
Demarcq, H., 2005. Modeling the primary and secondary productions of
the southern Benguela upwelling system: A comparative study through
two biogeochemical models. Global Biogeochemical Cycles 19, GB4021.
doi:10.1029/2004GB002427.
Large, W. G., McWilliams, J. C. & Doney, S. C.,1994. Oceanic vertical
mixing: a review and a model with a nonlocal boundary layer
parametrization. Reviews in Geophysics 32, 363-403.
Lazar, A., Vintzileos, A., Doblas-Reyes, F., Rogel, P. & Delécluse, P.,2005.
Seasonal forecast of tropical climate with coupled ocean-atmosphere
GCMs: On the respective role of the atmosphere and the ocean model
components in the drifting mean climate. Tellus 57, 387-397.
Lee, S.,K. and Csanady, G.,T., 1999a. Warm water formation and escape in
the upper tropical Atlantic Ocean: 1. A literature review. Journal of
Geophysical Research 104, 29561-29571.
Lee, S.,K. and Csanady, G.,T., 1999b. Warm water formation and escape in
the upper tropical Atlantic Ocean: 2. A numerical model study. Journal
of Geophysical Research 104, 29573-29590.
Lemos, C. F., Calbete, N. O., 1996. Sistemas que atuaram no Brasil de 1987
a 1995. Climanálise Especial, Edição comemorativa de 10 anos.
MCT/INPE-CPTEC.
Letelier, R. M., Dore, J. E., Winn, C. D., Karl, D. M., 1996. Seasonal and
interannual variations in photosynthetic carbon assimilation at Station
ALOHA. Deep-Sea Research II 43, 467-490.
Lett, C., Verley, P., Mullon, C., Parada, C., Brochier, T., Penven, P., Blanke,
B., 2008. A Lagrangian tool for modelling ichthyoplankton dynamics.
Environmental Modelling & Software 23, 1210-1214.
Liu, W. T., 2002. Progress in scatterometer applications. Journal of
Oceanography 58, 121-136.
Lumpkin, R., Speer, K., 2003. Large-scale vertical and horizontal circulation
114
in the North Atlantic ocean. Journal of Physical Oceanography 33, 1902-
1920.
Lumpkin, R., Garzoli, S. L., 2005. Near-surface circulation in the tropical
Atlantic Ocean. Deep Sea Research 52, 495–518.
Lutjeharms, J. R. E., Penven, P., Roy, C., 2003. Modelling the shear edge
eddies of the southern Agulhas Current. Continental Shelf Research 23,
1099-1115.
Malanotte-Rizzoli, P., Hedström, K., Arango, H. G., Haidvogel, D. B., 2000.
Water mass pathways between the subtropical and tropical ocean in a
climatological simulation of the North Atlantic. Dynamics of Atmosphere
and Oceans 32, 331-371.
Malone, T. C., Pike, S. E., Conley, D. J., 1993. Transient variations in
phytoplankton productivity at the JGOFS Bermuda time series station.
Deep-Sea Res. I, 40, 903-924
Marañón, E., Holligan, P. M.; Varela, M., 2000. Basin-scale variability of
phytoplankton biomass, production and growth in the Atlantic Ocean.
Deep-Sea Res. I, v. 47, p. 825-857.
Marchesiello, P., McWilliams, J. C., Shchepetkin, A., 2001. Open boundary
conditions for long-term integration of regional oceanic models. Ocean
Modelling 3, 1-20.
Marchesiello, P., McWilliams, J. C. & Shchepetkin, A., 2003. Equilibrium
structure and dynamics of the California Current System. Journal of
Physical Oceanography 33, 753-783.
Marchesiello, P., Lefevre, J., Penven, P., Lemarie, F., Debreu, L., Douillet, P.,
Vega, A., Derex, P., Echevin, V., Dewitte, B., 2008. Keys to affordable
regional marine forecast systems. La lettre trimestrielle Mercator Ocean
30, 38-48.
McCready, P., and Geyer, G.,R., 2001. Estuarine salt flux through an
isohaline surface. Journal of Geophysical Research 106, 11629-11637.
Meacham, S. P., 2000. Low-frequency variability in the wind-driven
circulation. Journal of Physical Oceanography 30, 269-293.
Messié, M., Radenac, M-H., 2006. Seasonal variability of the surface
chlorophyll in the western tropical Pacific from SeaWiFS data. Deep-Sea
115
Research I 53, 1581-1600.
Miller, J.R., 1976. The salinity effect ion a mixed layer ocean model. Journal
of Physical Oceanography 6, 29-35.
Miranda, L. B., Castro, B. M., 1981. Geostrophic flow conditions of the Brazil
Current at 19oS. Ciência Interamericana 22, 44-48.
Miyama, T., McCreary Jr., J. P., Sengupta, D., Senan, R., 2006. Dynamics of
biweekly oscillations in the equatorial Indian Ocean. Journal of Physical
Oceanography 36, 827-846.
Molinari, R., L., 1982. Observations of eastward currents in the tropical
South Atlantic Ocean: 1978-1980. Journal of Geophysical Research 87,
9707-9714.
Montes, M., J., F., 2003. Fatores que influenciam na produtividade dos
oceanos: A importância do fluxo de difusão dos nutrientes para a
biomassa do fitoplâncton na região oceânica do Nordeste brasileiro. PhD
Thesis Universidade Federal de Pernambuco, Recife, Brazil, 198 p..
Moore, J. K., Doney, S. C., Kleypas, J. A., Glover, D. M., Fung, I. Y., 2002.
An intermediate complexity marine ecosystem model for the global
domain. Deep-Sea Research II 49, 403–462.
Moura, A., Shukla, J., 1981. On the dynamics of droughts in Northeast
Brazil : Observations, theory, and numerical experiments with a general
circulation model. Journal of Atmospheric Science 38, 2653-2675.
Müller, T. J., Ikeda, Y., Zangenberg, N., Nonato, L. V., 1998. Direct
measurements of the western boundary currents between 20oS-28oS.
Journal of Geophysical Research 103, 529-543.
Mullon, C., Cury, P. & Penven, P. ,2002. Evolutionary individual-based
model for the recruitment of anchovy (Engraulis capensis) in the
southern Benguela. Canadian Journal of Fisheries.and Aquatic. Sciences
59, 910-922.
Murtugudde, R., Seager, R. & Busulacchi, A.,1996. Simulation of the
tropical oceans with an ocean GCM coupled to an atmospheric mixed-
layer model. Journal of Climate, 9:1795-1815.
Nobre, P., Campos, E., Polito, P., S., Sato, O.,T., and Lorenzzetti, J. (with
contribution by Robertson, I., Vianna, M., Servain, J., and Wainer, I.).
116
2004: PIRATA Western Extension Scientific Rational Report,
INPE/CPTEC Special Report, Cachoeira Paulista, SP, Brazil.
Nobre, P., Campos, E., Polito, P. S., Sato, O. T., Lorenzzetti, J. A., 2005.
Brazilian proposal for a PIRATA-SW Extension. Scientific Rationale. 44
pp.
Nobre, P., Campos, E., Servain, J. and Araujo, M.,2008. The PIRATA South-
West Extension: Scientific motivation, Realization and Perspectives (in
preparation).
OGCM-ECCO - Estimating the Circulation and Climate of the Ocean, 2007.
http://ecco.jpl.nasa.gov/, on line dataset, 12 September 2007.
Ohlmann, J.C., Siegel, D.A., Gautier, C., 1996. Ocean mixed layer depth
heating and solar penetration: A global analysis. Journal of Climate 9,
2265-2280.
Olson, D. B., Podesta, G. P., Evans, R. H., Brown, O. B., 1988. Temporal
variations in the separation of Brazil and Malvinas currents. Deep Sea
Research I 35, 1971-1990.
O’Reilly, J., E., Maritorena, S., Mitchell, B., G., Siegel, D., A., Carder, K, L.,
Garver, S., A., Kahru, M., McClain, C., 1998. Ocean color chlorophyll
algorithms for SeaWiFS. Journal of Geophysical Research, 103(C11),
24937-24953.
Penven, P., Roy, C., Colin de Verdière, A., Largier, J., 2000. Simulation and
quantification of a coastal jet retention process using a barotropic model.
Oceanological Acta 23, 615-634.
Penven, P., Roy, C., Lutjeharms, J. R. E., Colin de Verdière, A., Johnson, A.,
Shillington, F., Fréon, P., Brundrit, G., 2001a. A regional hydrodynamic
model of the Southern Benguala. South African Journal of Science 97,
472-476.
Penven, P., Lutjeharms, J. R. E., Marchesiello, P., Roy, C., Weeks, S. J.,
2001b. Generation of cyclonic eddies by the Agulhas Current in the lee
of the Agulhas Bank. Geophysical Research Letters 27, 1055-1058.
Penven, P., Debreu, L., Marchesiello, P., McWilliams, J. C., 2006. Application
of the ROMS embedding procedure for the Central California Upwelling
System. Ocean Modelling 12, 157-187.
117
Penven P., Marchesiello P., Debreu L., Lefevre J., 2008. Software tools for
pre- and post-processing of oceanic regional simulations. Environmental
Modelling and Software 23(5), 660-662.
Peterson R. G., Stramma, L., 1991. Upper-level circulation in the South
Atlantic ocean. Progress in Oceanography 26, 1-73.
Polito, P. S., Cornillon, P., 1997. Long baroclinic Rossby waves detected by
TOPEX/POSEIDON, Journal of Geophysical. Research 102, 3215–3235.
Proehl, J., 1996. Linear instability of equatorial zonal flows. Journal of
Physical Oceanography 26, 601–621
Rao, V. B., de Lima, M. C., Franchito, S. H., 1993. Seasonal and inter-
annual variations of rainfall over eastern Northeast Brazil. Journal of
Climate 6, 1754-1763.
Richardson, P., L., and McKee, T., K., 1984. Average Seasonal Variation of
the Atlantic Equatorial Currents from Historical Ship Drifts. Journal of
Physical Oceanography 14, 1226-1238.
Rodrigues, R. R., Rothstein, L. M., Wimbush, M., 2007. Seasonal variability
of the South Equatorial Current bifurcation in the Atlantic ocean: A
numerical study. Journal of Physical Oceanography 37, 16-37.
Schmid, C., Schafer, H., Podesta, G., Zenk, W., 1995. The Vitoria eddy and
its relation to the Brazil Current. Journal of Physical Oceanography 25,
2532-2546.
Schott, F. A., Stramma, L., Fischer, J., 1995. The warm water inflow into the
western tropical Atlantic boundary regime, spring 1994, Journal of
Geophysical. Research 100, 24745–24760.
Schott, F. A., Fischer, J., Stramma, L., 1998. Transports and pathways of
the upper-layer circulation in the western tropical Atlantic. Journal of
Physical Oceanography 28, 1904-1928.
Schott, F. A., Brandt, P., Hamann, M., Fischer, J., Stramma, L., 2002. On
the boundary flow off Brazil at 5-10oS and its connection to the interior
tropical Atlantic. Journal of Geophysical Research 29. doi:
10.1029/2002GL014786.
Schott, F., A., Dengler, M., Brandt, P., Affler, K., Fischer, J., Bourlès, B.,
Gouriou, Y., Molinari, R.,L., and Rhein, M., 2003. The zonal currents
118
and transports at 35°W in the tropical Atlantic. Geophysical Research
Letters 30, 1349. doi:10.1029/2002GL016849
Schott, F. A., Dengler, M., Zantropp, R., Stramma, L., Fischer, J., Brandt, P.,
2005. The shallow and deep western boundary circulation of the South
Atlantic at 5o – 11o S. Journal of Physical Oceanography 35, 2031–2053.
Schuchmann, K., 2006. Intraseasonal variability in the southwestern and
central tropical Atlantic ocean. PhD Thesis Christian-Albrechts-
Universität, Kiel, Germany, 148 p..
Servain J., Busalacchi, A., Moura, A., McPhaden, A., Reverdin, G., Vianna,
M., Zebiak, S., 1998. A Pilot Research Moored Array in the Tropical
Atlantic (PIRATA). Bulletin of the American Meteorological Society 79,
2019-2031.
Shchepetkin, A. F., McWilliams, J. C., 1998. Quasi-monotone advection
schemes based on explicit locally adaptive dissipation. Monthly Weather
Review 126, 1541-1580.
Shchepetkin, A. F., McWilliams, J. C., 2005, The regional oceanic modeling
system (ROMS): a split-explicit, free-surface, topography-following-
coordinate oceanic model, Ocean Modelling 9, 347-404.
She, J., Klinck, J. M., 2000. Flow near submarine canyons driven by
constant wind. Journal of Geophysical Research 105, 28671-28694.
Silveira, I. C. A., Miranda, L. B., Brown, W. S., 1994. On the origins of the
North Brazil Current. Journal of Geophysical Research 99, 22501-22512.
Smith, W. H. F., Sandwell, D. T., 1997. Global seafloor topography from
satellite altimetry and ship depth soundings. Science 277, 1957-1962.
Spall, M.A., 1991. A diagnostic study of wind- and buoyancy- driven north
Atlantic circulation. Journal of Geophysical Research 96, 18509-18518.
Sprintall, J., Tomczak, M., 1990. Salinity considerations in the oceanic
surface mixed layer. Ocean Sciences Institute Rep. 36, University of
Sidney, 170 pp.
Sprintall, J., Tomczak, M., 1992. Evidences of the barrier layer in the surface
layer of the tropics. Journal of Geophysical Research 97, 7305-7316.
Steeman-Nielsen, E., Jensen, E. A., 1957. Primary oceanic production, the
autotrophic production of organic matter in the oceans. Galathea Report
119
1, 49-136.
Stramma, L., Peterson, R. G., 1990. The South Atlantic Current. Journal of
Physical Oceanography 20, 846–859.
Stramma, L., 1991. Geostrophic transport of the South Equatorial Current
in the Atlantic. Journal of Marine Research 49, 281-284.
Stramma, L., Fischer, J. and Reppin, J., 1995. The North Brazil
Undercurrent. Deep Sea Research I 42, 773-395.
Stramma, L., England, M., 1999. On the water masses and mean circulation
of the South Atlantic Ocean. Journal of Geophysical Research 104,
20863–20883.
Stramma, L., Schott, F., 1999. The mean flow field of the tropical Atlantic
Ocean. Deep Sea Research 46B, 279–303.
Stramma, L., Fischer, J., Brandt, P., Schott, F., 2003. Circulation, variability
and near-equatorial meridional flow in the central tropical Atlantic. In:
Interhemispheric Water Exchanges in the Atlantic Ocean, G. J. Goni and
P. Malanotte-Rizzoli, editors, Elsevier. B.V., pp 1-22.
Stramma, L., Rhein, M., Brandt, P., Dengler, M., Böning, C., Walter, M.,
2005. Upper ocean circulation in the western tropical Atlantic in boreal
fall 2000. Deep Sea Research 52, 221–240.
Talley, L. D., 2003. Shallow, intermediate, and deep overturning components
of the global heat budget. Journal of Physical Oceanography 33, 530-560.
Thomas, W. H., 1970. On nitrogen deficiency in tropical Pacific Ocean
phytoplankton: photosynthetic parameters in rich and poor waters.
Limnology and Oceanography 15, 380-385.
Tsai, W. T., Spencer, M., Wu, C., Winn, C., Kellogg, K., 2000. SeaWinds on
QuikSCAT: sensor description and mission overview. Geoscience and
Remote Sensing ;symposium, 2000. Proceedings. IGARSS 2000. IEEE
2000 International, vol.3, no., pp. 1021-1023 vol. 3, 2000.
Veleda, D. R. A., 2008. Seasonal and intraseasonal variability of the western
boundary regime off the Eastern Brazilian Coast. PhD Thesis
Universidade Federal de Pernambuco, Recife, Brazil, 198 p..
Vintzileos, A., Delécluse, P. & Sadourny, R.,1999a. On the mechanisms in a
tropical ocean-global atmosphere coupled general circulation model. Part
120
121
I: Mean state and the seasonal cycle. Climate Dynamics, 15, 43–62.
Vintzileos, A., Delécluse, P., Sadourny, R.,1999b. On the mechanisms in a
tropical ocean-global atmosphere coupled general circulation model. Part
II: Inter-annual variability and its relation to the seasonal cycle. Climate
Dynamics, 15, 63–80.
Weisberg, R., H., 1985. Equatorial Atlantic velocity and temperature
observations: February-November 1981. Journal of Physical
Oceanography 15, 533-543.
Wienders, N., Arhan, M., Mercier, H., 2000. Circulation at the western
boundary of the south and equatorial Atlantic. Journal of Marine
Research 58, 1007-1039.
Woodruff, S., Diaz, H. F., Elms, J. D., Workey, S. J., 1998. COADS release 2
data and metadata enhancements for improvements of Marine Surface
Flux Fields. Physics and Chemistry of the Earth 23,517-526.
Zhang, H. M., Bates, J. J., Reynolds, R. W., 2006. Assessment of composite
global sampling: sea surface wind speed. Geophysical Research Letters
33, L17714: 1-5.
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo