343
TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade de Ciencias Universidade de Vigo Spatio-temporal variability of the physical and biogeochemical conditions in the middle Ría de Vigo’ Memoria presentada por la Lcda. en Ciencias del Mar Silvia Piedracoba Varela para optar al título de Doctor por la Universidad de Vigo Vigo, octubre de 2006 Variabilidad espacio-temporal de las condiciones físicas y químicas en un segmento central de la Ría de Vigo’ UNIVERSIDADE DE VIGO

digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade de Ciencias Universidade de Vigo

‘Spatio-temporal variability of

the physical and biogeochemical conditions in the middle

Ría de Vigo’

Memoria presentada por la Lcda. en Ciencias del Mar Silvia Piedracoba Varela para optar al título de Doctor por la Universidad de Vigo Vigo, octubre de 2006

‘Variabilidad espacio-temporal de las condiciones físicas y químicas

en un segmento central de la Ría de Vigo’

UNIVERSIDADE DE VIGO

Page 2: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 3: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

El Dr. Gabriel Rosón Porto, Profesor Titular de la Universidad de Vigo, y

el Dr. Xosé Antón Álvarez Salgado, Investigador Científico, del Consejo

Superior de Investigaciones Científicas (CSIC),

HACEN CONSTAR: Que la presente memoria, titulada

“Variabilidad espacio-temporal de las condiciones físicas y

químicas en un segmento central de la Ría de Vigo / Spatio-

temporal variability of the physical and biogeochemical conditions in the

middle Ría de Vigo”, presentada por

la licenciada en Ciencias del Mar Silvia Piedracoba Varela para optar al

grado de Doctor por la Universidad de Vigo, ha sido realizada bajo

nuestra dirección, cumpliendo las condiciones exigidas para su

presentación, la cual autorizamos.

Y para que así conste y surta los efectos oportunos, firmamos la presente

en Vigo, a 27 de septiembre de 2006.

Fdo.: G. Rosón Porto Fdo.: X. A. Álvarez Salgado

UNIVERSIDADE DE VIGO

Page 4: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 5: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

D. Miguel Ángel Nombela Castaño, Profesor Titular del Departamento de

Xeociencias Mariñas e Ordenación do Territorio de la Universidad de

Vigo, y Tutor de la presente memoria de Tesis Doctoral titulada:

“Variabilidad espacio-temporal de las condiciones físicas y químicas en

un segmento central de la Ría de Vigo/Spatio-temporal variability of the

physical and biogeochemical conditions in the middle Ría de Vigo”, de la que es

autora Silvia Piedracoba Varela y que ha sido dirigida por el Dr. Gabriel

Rosón Porto, Profesor Titular de la Universidad de Vigo y el Dr. Xosé

Antón Álvarez Salgado, Investigador Científico del Consejo Superior de

Investigaciones Científicas (CSIC),

AUTORIZA su presentación en la Universidad de Vigo para su

lectura y defensa.

Y para que así conste y surta los efectos oportunos, firmo la presente

en Vigo, a 27 de septiembre de 2006.

Fdo.: Miguel Ángel Nombela Castaño

UNIVERSIDADE DE VIGO

Page 6: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 7: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Esta Tesis fue realizada durante el disfrute de una beca de Formación

de Personal Investigador del Ministerio de Educación y Ciencia en la

Universidad de Vigo y una beca de postgrado del programa I3P del

Consejo Superior de Investigaciones Científicas.

Los trabajos que se recogen en esta memoria fueron financiados por el

proyecto del Programa Nacional de Ciencias y Tecnologías Marinas

REactividad de la Materia Orgánica Disuelta en el Sistema de

Afloramiento de la Ría de Vigo REMODA (CICYT-REN-2000-0800-C02-

01-02) e incentivados por la Xunta de Galicia (PGIDT01MAR40201PN).

Además, en el transcurso de la elaboración de esta Tesis Doctoral, se

realizó una estancia de dos meses en el School of Ocean Science (Bangor,

Wales, UK) financiada por el Ministerio de Educación y Ciencia.

Page 8: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 9: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Agradecimientos-Acknowledgements

Bueno ahí voy… Más de uno ya estará sonriendo pensando: ‘Ahí va la

embalada’. Aunque esto de los agradecimientos me sabe mejor hacerlo en

persona, lo bueno de escribirlos es que siempre quedarán plasmados y

que es la antesala más humana y personal de lo que viene a continuación.

Aquí se supone que puedo escribir lo que quiera y que no tengo

revisores….o sí, pero en plan cotilla ;-)

No me puedo creer que haya llegado aquí. Siempre he dicho que una

Tesis es como parir un hijo lentamente. Y eso que yo aún no sé lo que es

eso!!! Bueno, pues esto es un parto multitudinario, por la cantidad de

personas que han colaborado. He tenido la suerte de poder realizar esta

Tesis bajo la dirección de dos grandes personas, Gabi y Pepe, y en dos

laboratorios en los que me he sentido increíblemente a gusto. Gabi,

siempre me has facilitado las cosas, resuelto dudas, me has dado buenos

consejos y una visión objetiva cuando me iba por las ramas a lo largo de

todos estos años. Pepe, es cierto que eres un jefe muy paternal, siempre

preocupado porque todo salga adelante, y lo mejor de todo, transmitiendo

el entusiasmo que sientes por la Ciencia. He aprendido muchísimo

contigo. Más de lo que pudiera imaginar.

Mi etapa en el Laboratorio de Física fue la fase de partida, donde

empecé a arrancar. Hubo muchos momentos en que pensaba: ¿Será esto lo

mío? Pero supongo que esas dudas son normales, son dudas de

principiante. Y lo que es cierto, es que Gabi y Ramiro han tenido mucho

que ver en que tuviera más confianza en mí misma. Ramiro, no es que

seas mi Director de Tesis pero el hecho de que fueras el Investigador

Principal de REMODA implicó que trabajásemos juntos, y que

Page 10: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

compartiéramos horas de barco y de análisis de datos. A lo largo de las

campañas de Remoda pasé de decir esto es demasiado para mí a sentirme

cómoda, y en eso has tenido mucho que ver tú. Los demás Gofuvis

(aunque ahora ya no estéis todos): Juan, Carlos, Luís, Belén, Silvia, Ángel,

Rocío, Nacho, Pablo…miles de gracias por vuestro tiempo, tanto a nivel

científico como personal. Algunos habéis sido mis maestros cuando

empezaba y eso facilita mucho el poder avanzar.

Cuando llegué al Laboratorio de Oceanoloxía todos los que allí estabais,

con vuestra acogida, me hicisteis sentir muy a gusto. Paula, Bibiana, Mar,

Samantha, Toni, Fer, Miguel, Suso, Isa, Mónica, Vane, Pili, Rosa, Chus,

Rita, Marcos, Óscar, Belén, Luisa, Roberta, Marta, miles de gracias.

Muchos de vosotros formáis parte de REMODA y este trabajo es también

vuestro. Esta Tesis es el resultado de muchos análisis que han llevado su

tiempo de laboratorio. Si no fuera por vosotros no sería posible.

Y a la cúpula del laboratorio de Oceanoloxía (pero que no requiere

protocolo porque son todos muy cercanos): Aída, Carmen, Paco, Des, Fiz,

gracias por vuestros útiles consejos y por haberme echado un cable

siempre que os le he pedido.

Tanto en un laboratorio como en el otro muchos de vosotros no sólo

fuisteis o sois compis de trabajo, sino que además comparto una amistad.

Ha habido muchas etapas en el desarrollo de esta Tesis, porque al fin y a

al cabo 5 añitos dan mucho de si. En unas he sido más feliz que en otras

porque como ser humano que soy una no es de piedra... Digamos que soy

“desacorde” con mi apellido. Pero lo cierto es que me siento muy

afortunada porque siempre tuve donde apoyarme.

La verdad es que nada sería lo mismo sin esas comidas compartidas

entre Gofuvis y la gente tan maja del Laboratorio de Ecología en el Cuvi.

Algunos no están, otros han vuelto y también hay nuevas incorporaciones.

Page 11: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Es que es un laboratorio con mucha circulación. Patri, Vales, Patri E, Gon,

Nacho, Mai, Nuri, Juan B, Esther, Bea, Pedro, Eva, María A, María E, Jose,

Lili, Cris, Carmen, Paula, Sandra, Ana… Graciñas por esos momentos de

distensión en los que me he reído mucho, ya sea en el Cuvi o de campaña.

Tampoco nada sería igual sin los cafés en el Tertulia o las comidas en

Bouzas, sobre todo en la Leman… Me quedan muy buenos recuerdos de

ambas etapas.

Y por el camino 2 mesecitos en Bangor. Thank you very much to the

people of SOS: John, Tom, Phil, Catherine, Matt, Neil, Joao. I felt very

received with you.

Esos dos meses en Bangor no serían lo mismo sin la tropa con la que

vivía. Miriam, Vitish, Elba, Max, Toni, Enre… todavía recuerdo cómo

convencisteis a aquel “propietario” de que tenía que vivir allí, y también

recuerdo con tristeza el día que me despedí de vosotros. Me supo a poco

mi tiempo allí.

Esta Tesis lleva detrás muchas horas de barco con una tripulación que

para mi siempre será la tripulación del Mytilus. Así es como os veo yo.

Jorge, Apo, Pirri, Waldo, Ricardo, los días de muestreo no serían lo mismo

sin vosotros.

Quiero nombrar a tres personas de una forma especial porque habéis

estado “siempre” ahí todos estos años, y porque hemos compartido

mucho, y porque tengo la suerte de que seguís a mi lado, y porque, y

porque… Patri, Pau, Silvia, creo que no me llega la palabra GRACIAS

para vosotras…

Elena, David, Vero, Lucía, Natalia, Maria B, María A, Luisi, Luci,

Carola, Santi, millones de gracias por vuestro apoyo y por vuestra

amistad!!! Ojalá os lo pudiera decir a todos en persona… Hay algunas

personas que en su día fueron muy importantes y ahora por

Page 12: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

circunstancias no forman parte de mi vida. Supongo que a la hora de

agradecer no me tengo que centrar en la etapa final de algo tan largo como

una Tesis.

Probablemente me quede gente por el camino. Espero que no se pique

si llega a leer esto y no se encuentra. A lo mejor me olvido en un papel

pero seguro que en persona no, que al fin y al cabo es lo que importa.

Dejo para el final a mi familia..., aunque sea lo primero para mí. Y en la

familia me quiero centrar en mis padres y en mi hermano. Si no fuera por

vosotros, papá y mamá, yo no habría llegado aquí. Y no me refiero

precisamente a que me hayáis dado la vida sino a lo que ha sido mi vida

por teneros a vosotros a mi lado. Sabemos que hubo momentos difíciles

pero también nos hemos recuperado y eso es lo que cuenta al final.

Además…, si no fuera por los momentos chungos no valoraríamos lo

bueno y nuestra vida sería muy lineal y muy aburrida, o no?

Y a ti Juan, creo que si me dieran la posibilidad de cambiar de hermano

jamás lo haría ;-) Es más… me han compensado las tortas que recibí de

pequeña por tenerte no solo como hermano sino como amigo.

Y aprovecho la parte sensible y sentimental de la Tesis para terminar

con una frase que no es mía pero que me encanta:

"¿Por qué nos gusta el mar? Es porque tiene una poderosa capacidad

para hacernos pensar cosas que nos gusta pensar."

Robert Henri (Pintor estadounidense, 1865-1929)

Creo que conmigo lo consigue.

Page 13: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

I

Chapter 1

Introduction.……………………………………………………… 5

1. El sistema de afloramiento del NO de España.……………. 7

2. Geomorfología de las rías ……………………………………. 18

3. Meteorología de las Rías Baixas …………………………….. . 22

4. Circulación de las Rías Baixas………………………………….. 28

5. Hidrodinámica y biogeoquímica de las Rías Baixas .………. 32

5.1. Conocimiento hidrodinámico previo al proyecto

REMODA………………………………………………………. 33

5.2. Conocimiento biogeoquímico previo al proyecto

REMODA………………………………………….…………. 41

5.2.1. Impacto de la hidrografía y la dinámica en los ciclos

biológicos y geoquímicos de las Rías Baixas …….……. 41

5.2.2. Balance biogeoquímico de las rías ……………………….. 49

6. Hipótesis general .………….…………………………………… 61

7. Objetivos …………………………………………………… …… 61

8. Estructura ………………………………………………………… 63

Chapter 2

Short-timescale thermohaline variability and residual circulation in the central segment of the coastal upwelling system of the Ría de Vigo …………………………………………………………………. 65

Resumen …………………………………………………………….. 67

Abstract …………………………………………………………….. 69

1. Introduction ……………………………………………………... 71

2. Material and methods ………………………………………….. 75

2.1. Measured Variables …………………………………………… 75

2.2. Variables calculated from collected data …………………… 77

2.3. Estimation of water flows with an inverse method ……….. 79

Page 14: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

II

2.4. Empirical orthogonal functions (EOF) analysis of

local winds ………………………………………………………… 82

2.5. Cross correlation between shelf wind and residual

currents ……………………………………………………………. 83

3. Results ……………………………………………………………….. 83

3.1. Local and remote winds ………………………………………….. 83

3.2. Residual currents ………………………………………………….. 86

3.3. Thermohaline variability …………………………………………. 94

4. Discussion …………………………………………………………… 101

4.1. Remote winds, a key forcing agent of the

subtidal motion in estuaries and coastal inlets ………………… 102

4.2. Coupling between hydrography and

hydrodynamics ……………………………………………………. 107

5. Conclusions …………………………………………………………. 113

Chapter 3

Lateral variability in the coastal upwelling system of the

Ría de Vigo …………………………………….……………………… 115

Resumen ……………………………………………………………….... 117

Abstract …………………………………………………………………. 119

1. Introduction …………………………………………………………. 121

2. Material and methods ……………………………………………… 123

2.1. Study site …………………………………………………………… 123

2.2. Data collection …………………………………………………….. 124

2.3. Limitations…………………………………………………………. 129

2.4. Processing and accuracy …………………………………………. 129

2.5. Quality control …………………………………………………….. 131

2.6. Residual versus total current …………………………………….. 131

2.7. Numerical model ………………………………………………….. 132

Page 15: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

III

3. Results ……………………………………………………………….. 135

3.1. Morphology of the ría …………………………………………….. 135

3.2 Remote and local winds …………………………………………... 136

3.3 Residual currents …………………………………………………... 137

3.4. Salinity and temperature …………………………………………. 141

3.5. Level of no movement, halocline and

termocline …………………………………………………………... 145

4. Discussion …………………………………………………………… 146

4.1. Factors controlling the lateral circulation

in coastal embayments and estuaries ……………………………. 146

4.2. Numerical model outputs versus field

observations ………………………………………………………... 153

4.3. Importance of the lateral circulation in

coastal embayments and estuaries ………………………………. 157

5. Conclusions ………………………………………………………….. 159

Chapter 4

Origin and fate of a bloom of Skeletonema costatum

during a winter upwelling/downwelling sequence in

the Ría de Vigo ……………………………………………………….. 161

Resumen ………………………………………………………………… 163

Abstract ………………………………………………………………….. 165

1. Introduction …………………………………………………………. 167

2. Material and Methods……………………………………………... 170

3. Results ………………………………………………………………... 179

3.1. Water transports ………………………………………………….... 179

Page 16: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

IV

3.2. Biochemical composition of the materials produced

during the bloom ………………………………………………….. 179

3.3. Production and consumption rates of Chla during

the course of the bloom …………………………………………… 186

4. Discussion …………………………………………………………… 189

5. Conclusions ………………………………………………………….. 197

Chapter 5

Physical-biological coupling in the coastal upwelling system

of the Ría de Vigo: An in situ approach ……………………..……... 199

Resumen .……………………………………………………………….. 201

Abstract .………………………………………………………………... 203

1. Introduction …………………………………………………………. 205

2. Material and Methods ……………………………………………… 207

3. Results ……………………………………………………………….. 212

3.1. Water flows ………………………………………………………… 212

3.2. Nitrogen fluxes …………………………………………….............. 224

3.3. Nitrogen budget …………………………………………………… 232

4. Discussion ………………………………………………………….. 236

4.1. Role of vertical convection and diffusion in the fertilisation

of the photic layer ………………………………………………... 236

4.2. A geochemical approach to the origin and fate of biogenic

materials .………………………………………………………….. 240

5. Summary ……………………………………………………………. 244

Chapter 6

Physical-biological coupling in the coastal upwelling system

of the Ría de Vigo. An in vitro approach …………………….…… 245

Resumen .………………………………………………………………. 247

Page 17: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

V

Abstract ……………………………………………………………….... 249

1. Introduction ……………………………………………………….... 251

2. Material and methods ……………………………………………... 252

3. Results ……………………………………………………………….. 256

3.1. Chemical composition of the biogenic materials ……………… 256

3.2. Metabolic balance of the microbial loop in relation to

phytoplankton population ………………………………………. 259

3.3. Microzooplankton grazing ……………………………………….. 265

3.4. Vertical fluxes of biogenic organic materials …………………… 267

4. Discussion …………………………………………………………… 269

5. Conclusions …………………………………………………………. 279

Chapter 7

General Conclusions …………………………………………………. 281

Chapter 8

Literature Cited ……………………………………………………….. 291

Page 18: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Table of contents

VI

Page 19: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Glossary

1

Glossary of relevant terms A Free surface area of the ría (in m2) AS, AB Surface and bottom areas of the open boundary (in m2) Ah, Az Horizontal and vertical eddy viscosities (in m2/s) ADCP Acoustic Doppler Current Meter α Coefficient of thermal expansion (in ºC-1) β Coefficient of haline contraction (in psu-1) B Heat exchange with the atmosphere by back radiation (in cal cm–2 d–1) BT Bottom tracking Chl0 Initial chlorophyll concentration (in mg m-3) Chl a Chlorophyll a (in mg m-3) <Chl> Average Chl concentration in the photic layer (in mg m-3) C Heat exchange with the atmosphere by conduction (in cal cm–2 d–1) CD Empirical drag coefficient (dimensionless) CS, CB Average surface and bottom concentrations of the specie C C0, Ct Initial and final carbon biomass concentrations Cho Particulate carbohydrates (in µmol C L–1) d–Cho Total dissolved carbohydrates (in µmol C L–1) D Dilution factor of the sample Dep Deposition fluxes (in g N m-2 d-1 or g C m-2 d-1) DOC, DON, Dissolved organic carbon, nitrogen and phosphorous DOP (in µmol L–1)

GPSδ Error introduced in SVr

by the GPS (in cm s-1) es Vapour pressure the sea surface (in mbar) ez Vapour pressure at 2 m above the sea (in mbar)

ATe Distilled water vapour pressure at AT (in mbar)

STe Distilled water vapour pressure at ST (in mbar) E Heat exchange with the atmosphere by evaporation (in cal cm–2 d–1) ENACW Eastern North Atlantic Central Water ENACWp ENACW of subpolar origin ENACWt ENACW of subtropical origin EOF Empirical Orthogonal Functions fC Coriolis parameter (in s-1) Fx Flux of any species carried by a horizontal convective flow (QS, QB)

Page 20: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Glossary

2

ZF Flux of any species carried by vertical convective flow (QZ) g Instantaneous microzooplankton grazing rate (in d-1) G Grazing rate (in g m-2 d-1 of N/C) GPS Global Positioning System H Heat exchange with the atmosphere (in cal cm–2 d–1) HPLC High Performance Liquid Chromatography Hu Relative humidity I Irradiation (in cal cm–2 d–1) ĸ Daily net growth rate (in d-1) k Von Karman’s constant (dimensionless) Kz Turbulent mixing coefficient (in m2 d-1) Lip Particulate lipids LNM Level of no movement MCho, Dissolved mono– and polysaccharides PCho µ Instantaneous rates of growth (in d-1) <µ> Average µ in the photic layer (in d-1)

ZM Vertical diffusive fluxes N Cloudiness <N> Brunt-Väsisälä frequency (s-1) NT Dissolved inorganic nitrogen (in µmol L–1) NCP Net community production (in mmol O2 m-2d-1 or g m-2d-1 of N/C/ Chl) NEP Net ecosystem production (in g m-2 d-1 of N/C) NEPT Net ecosystem production of NT (in g N m-2d-1) NEPD Net ecosystem production of DON (in g N m-2d-1)

NEPP Net ecosystem production of PON (in g N m-2d-1)

NEPSED Net balance of PON sedimentation, PON and DON resuspension and NT diffusion from the sediment (in g N m-2d-1) NT Neap tide o – i Net budget of outputs minus inputs of any chemical species (o – i)T Net budget of outputs minus inputs of NT (in g N m-2d-1) (o – i)P Net budget of outputs minus inputs of PON (in g N m-2d-1)

Page 21: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Glossary

3

(o – i)D Net budget of outputs minus inputs of DON (in g N m-2d-1) P Precipitation (in mm d-1) PAR Photosynthetically available radiation Pg Gross primary production or daily photosynthetic rate Pho Phosphorous compounds POC, PON, Particulate organic carbon, nitrogen and phosphorous POP POM Particulate organic matter Pr Atmospheric pressure (in mbar) Prt Particulate proteins QB Bottom horizontal flow QS Surface horizontal flow QZ Vertical flow R Reflection R Respiration rate (in mmol O2 m-2d-1 or g m-2d-1 of N/C/ Chl) Rc, Rn, Rp Stoichiometric ratios of conversion of dissolved oxygen production into organic carbon, nitrogen, and phosphorus Ri Richardson number ρ Density ρair Density of air ρSW Density of seawater SD Standard deviation Sed Sedimentation rate (in g m-2 d-1 of N/C) ST Spring Tide TA, TS Air and water surface temperature (in ºC) TDN Total dissolved nitrogen (in µmol N L–1) TOC Total organic carbon (in µmol C L–1) u* Friction velocity (in cm s-1)

0U Tidal velocity amplitude (in cm s-1) V Volume (in m3)

CVr

Current velocity (in cm s-1)

SVr

Ship velocity (in cm s-1)

TVr

Total velocity measured by the VMADCP (in cm s-1) VZ Vertical convective velocity (in cm s-1) VMADCP Vessel Mounted Acoustic Doppler Current Profiler W Wind speed (in m s-1)

Page 22: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Glossary

4

ϖ Sedimentation velocity (m d-1) ∆(V·NT)/ ∆t Accumulation of NT (in g N m-2d-1)

∆(V·DON)/ ∆t Accumulation of DON (in g N m-2d-1)

∆(V·PON)/ ∆t Accumulation of PON (in g N m-2d-1)

<µ> Average µ in the photic layer (in d-1)

Page 23: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1:

Introduction

Page 24: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

     

Page 25: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

7

Esta Tesis se enmarca en el proyecto “REMODA” (REactividad de la

Materia Orgánica Disuelta en el Sistema de Afloramiento de la Ría de

Vigo), referencias REN2000–0880–C02–01 y –02 MAR, que se solicitó en la

convocatoria del Plan Nacional de I+D+I del del año 2000 y se ejecutó

entre los años 2001 y 2003. Por otra parte, esta Tesis es la contribución nº

38 de la Unidad Asociada GOFUVI–CSIC.

Esta memoria consta de cinco capítulos independientes, cada uno de

los cuales tiene su propia introducción. Por lo tanto, este primer capítulo

tiene como objetivo presentar las características geomorfológicas,

meteorológicas y dinámicas de las Rías Baixas en general, haciendo énfasis

en el estado de conocimiento de los procesos hidrodinámicos,

biogeoquímicos y del acoplamiento entre ambos, en el momento en el que

se solicitó el proyecto. Finalmente, se presentarán la hipótesis, los

objetivos específicos y la estructura de la memoria.

1. El sistema de afloramiento del NO de España

Las Rías Baixas en general, y la Ría de Vigo en particular, se encuentran

en el límite septentrional de uno de los ecosistemas marinos más

productivos del mundo: el sistema de afloramiento costero del NO de

Africa/Península Ibérica (Figura 1), que se extiende entre las latitudes

10ºN y 44ºN y es uno de los cuatro grandes sistemas de afloramiento

costero junto con California/Oregón, Perú/Chile y Namibia/Sudáfrica

(Bakun y Nelson 1991).

A nuestras latitudes (42–44ºN), la circulación superficial se relaciona

con los cambios estacionales observados en el régimen de vientos sobre la

plataforma continental. Desde Marzo−Abril a Septiembre−Octubre, el

Anticiclón de las Azores se refuerza, situándose

Page 26: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

8

Figure 1. Mapa del margen Este del Atlántico Norte, sistema de afloramiento costero del NO de Africa/Península Ibérica. En detalle, las costas gallegas.

en el centro del Atlántico Norte al tiempo que la Borrasca de Islandia se

debilita. Durante este periodo dominan vientos de componentes Noroeste,

favorables al afloramiento (Wooster et al. 1976; Bakun y Nelson 1991),

dando lugar a una corriente superficial en la plataforma y el talud dirigida

hacia el Sur (Torres et al. 2003; Álvarez-Salgado et al. 2003). Por el

contrario, el resto del año prevalecen vientos de componente Sureste,

Page 27: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

9

favorables al hundimiento, a consecuencia del desplazamiento del

Anticiclón de las Azores hacia el Noroeste de África y el reforzamiento de

la Borrasca de Islandia (Wooster et al. 1976; Bakun y Nelson 1991;

Nogueira et al. 1997). La Figura 2 muestra el ciclo estacional de las

componentes Este y Norte del viento que sopla sobre la plataforma

continental frente a las Rías Baixas.

La masa de agua dominante en las rías gallegas y la plataforma

adyacente es ENACW (Eastern North Atlantic Central Water) o ACNAE

(Agua Central Noratlántica del Este) con dos ramales (Fiúza 1984; Ríos et

al. 1992; Catro 1997): 1) ENACWt (Eastern North Atlantic Central Water of

tropical origin) o ACNAEt (Agua Central Noratlántica del Este de origen

subtropical), cálida (>13ºC) y salada (>35,66) que se forma al Sur de 40º en

capas de <150m y es transportada hacia el Norte. 2) ENACWp (Eastern

North Atlantic Central Water of tropical origin) o ACNAEp (Agua Central

del Atlántico Noroeste de origen subpolar) que se forma durante el

invierno al norte de 43ºN en capas de hasta 600 m de profundidad,

presenta temperaturas <13ºC y es transportada hacia el Sur.

El ENACW, que se caracteriza por una relación lineal positiva entre

salinidad y temperatura (Figura 3a). Esta relación no es fija de año en año,

sino que cambia con una frecuencia decadal en función de condiciones

atmosféricas regionales, sujetas a la Oscilación del Atlántico Norte (Pérez

et al. 1995; 2000a). El ENACW lleva disueltas las sales nutrientes que el

fitoplancton de las rías precisa para crecer y dividirse.

Page 28: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

10

E F M A M J J A S O N D

V es

te (m

s-1)

-15

-10

-5

0

5

10

15

E F M A M J J A S O N D

V no

rte (m

s-1)

-15

-10

-5

0

5

10

15

1 14 27 40 53

V es

te y

V no

rte (m

s-1)

-8.0-6.0-4.0-2.00.02.04.06.08.0

Figura 2. Valores mensuales de las componentes Este–Oeste y Norte–Sur del viento que sopla sobre la plataforma frente a las Rías Baixas obtenidos a partir de una serie diaria de datos entre 1982 y 2000. Estos datos de velocidad geóstrofica en el 43ºN y 11ºW han sido calculados por Cabanas J.M. (Instituto Español de Oceanografía, Vigo). En las gráficas (a) y (b) se muestran la media (línea de puntos), mediana (línea continua), el rango en el que se encuentran el 50% de las observaciones (caja) y el rango en el que se encuentran el 80% de las observaciones (bigotes). En la gráfica c se muestran los valores medios semanales, reales (puntos) y suavizados (líneas) en el periodo 1982–2000 para las dos componentes del viento

a

b

c

Page 29: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

11

Longitud (W)-11 -10 -9 -8

Longitud (W)-11 -10 -9 -8

Latit

ud (N

)

42

43

44

Cabo Fisterra

Rio Miño

-14 -10 -6

-14 -10 -6

Latit

ud (N

)

30

35

40

45O

céan

o A

tlánt

ico

Golfo de Vizcaia

Pení

nsul

a Ib

éric

aA

fric

a

Rías Baixas

S

35.5 35.6 35.7 35.8 35.9 36.0

θ (ºC

)

10

11

12

13

14

15

16

a

θ (ºC)10 11 12 13 14 15 16

NO

3 (µm

ol k

g-1)

0

2

4

6

8

10

12

14

16

b

En la Figura 3b se muestra la relación lineal entre el nitrato y la

temperatura que se encuentra en la Cuenca Ibérica. La concentración de

nitrato aumenta al disminuir la temperatura, de modo que las aguas más

frías (ENACWp) tienen niveles de nutrientes mayores que aguas más

cálidas (ENACWt).

Figura 3. Relación entre salinidad y temperatura (a) y nitrato (NO3–) y temperatura (b) en el Agua Central del Atlántico Noroeste (ENACW). El mapa adjunto muestra las estaciones empleadas para establecer as relación entre salinidad, nitrato y temperatura. La línea que aparece en el diagrama temperatura/salinidad es la línea de referencia de ENACW, establecida por Fiúza (1984) y Ríos et al. (1992a). Tomado de Álvarez–Salgado et al. (2002).

Page 30: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

12

Consecuentemente, las diferencias espaciales y temporales en la

distribución de las distintas ramas de ENACW en las rías gallegas

afectarán igualmente a la fertilización de las mismas.

En océano abierto al Oeste de 10º30’ W (Mazé et al. 1997), la corriente

de Portugal o Ibérica (Portugal o Iberian current; PC o IC) transporta

ENACWp hacia el sur (Pollard y Pu 1985; Krauss 1986; Sy 1988; Pérez et

al. 2001; Martíns et al. 2002; Peliz et al. 2005) y al Este de 10º30’ W y al

Norte de 40ºN la Contracorriente Costera de Portugal (Portugal Coastal

Counter Current o Iberian Poleward Current; PCCC o IPC) trasporta

ENACWt hacia el Norte en la fachada atlántica (Pingree y Le Cann 1989;

Frouin et al. 1990; Haynes y Barton 1990) y hacia el Este en la fachada

cantábrica (Pingree y Le Cann 1990).

Desde un punto de vista general, podemos decir que este tipo de

circulación hacia el polo es habitual en el margen oriental de los grandes

océanos (Pingree 1994; McCreary et al. 1986; Frouin et al. 1990) y suele

estar forzada por gradientes geopotenciales de larga escala y por la

batimetría, constituyendo un mecanismo de compensación de la

circulación general hacia el Ecuador. Sin embargo, Frouin et al. (1990)

puntualizó que este flujo dirigido hacia el Norte se debe principalmente a

la compensación geopotencial, que influye con un 80%, pero también al

viento local, que influye con un 20%.

A lo largo de la costa gallega se forma un frente zonal que separa las

aguas tropicales y subpolares, que migra desde el Río Miño en la

Primavera y hasta el río Eo en Otoño (Castro 1997). Por lo tanto, hay

diferencias espaciales (entre rías) como temporales (dentro de una misma

ría) en la calidad del ENACW que entra en las rías gallegas.

Page 31: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

13

Durante la época de afloramiento, vientos del Noreste dan lugar a una

corriente superficial hacia el Sur sobre la plataforma y el talud (Torres et

al. 2003) de modo que la IPC se desplaza hacia océano abierto y/o fluye

como una contracorriente subsuperficial (Péliz et al. 2005; Figura 4). Los

vientos de componente Norte producen afloramiento de agua

subsuperficial desde 150–200 metros de profundidad. El fenómeno de

afloramiento también se detecta fácilmente en las imágenes de satélite

(Figura 4). En ella se observa un frente de afloramiento, con agua fría en la

plataforma frente a las Rías Baixas. El agua que aflora es ENACW. Según

transcurre el verano, el frente de afloramiento suele evolucionar hacia

estructuras más complejas como el centro de afloramiento del Cabo

Finisterre y el filamento de las Rías Baixas (Haynes et al. 1993), tal como

muestra la imagen de satélite (Figura 4). Estas estructuras juegan un papel

crucial en la evacuación de agua de la costa hacia el océano (Joint et al.

2001).

Sin embargo, en la época de hundimiento la IPC fluye desde la

superficie en la plataforma en respuesta a los vientos dominantes del

Suroeste (Figura 5). La presencia de agua de origen subtropical sobre la

plataforma ibérica en respuesta a vientos de componente Sur se detecta

en imágenes de satélite (Frouin et al. 1990; Haynes y Barton 1990; Pingree

y Le Cann 1990) como la de la Figura 5.

La cálida y salina ENACWt transportada por la IPC forma un frente de

densidad meridional próximo a la costa con los aportes continentales

salobres debidos a los ríos Verdugo–Oitabén, Lérez, Ulla, Umia y Tambre,

así como la contribución del Río Miño, que generan una pluma de agua

salobre que se extiende a lo largo de la costa y que se conoce como la

WIBP (Western Iberia Buoyant Pluma; Peliz et al. 2002). La WIBP

Page 32: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

14

frecuentemente se localiza en el talud, aunque su posición varía en

función del viento y de los aportes continentales.

Figura 4. Imagen de satélite AVHRR y patrón de circulación típico en Primavera y Verano. Colores “cálidos” en la imagen de satélite (amarillo, naranja y rojo) indican temperaturas progresivamente mayores. Por el contrario, colores “fríos” (verde, azul y morado) indican temperaturas progresivamente menores. Igualmente se muestra la circulación de agua, tanto en dirección paralela como perpendicular a la costa. Imágenes cedidas por S. Groom, Remote Sensing Group, Plymouth Marine Laboratory (Reino Unido). Esquemas de circulación tomados de Álvarez-Salgado y Fraga (en prensa).

Page 33: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

15

Figura 5. Imagen de satélite AVHRR y patrón de circulación típico en Otoño y Invierno. Colores “cálidos” en la imagen de satélite (amarillo, naranja y rojo) indican temperaturas progresivamente mayores. Por el contrario, colores “fríos” (verde, azul y morado) indican temperaturas progresivamente menores. Igualmente se muestra la circulación de agua, tanto en dirección paralela como perpendicular a la costa. Imágenes cedidas por S. Groom, Remote Sensing Group, Plymouth Marine Laboratory (Reino Unido). Esquemas de circulación tomados de Álvarez-Salgado y Fraga (en prensa).

Page 34: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

16

En la Figura 6 se muestra una sucesión de 12 imágenes de satélite

(medias mensuales de 2002, año de muestreo a partir del cual se

realizaron los trabajos correspondientes a esta memoria) de la temperatura

superficial del mar en las costas de Galicia y océano adyacente. En ellas se

puede observar la evolución desde las frías temperaturas superficiales

durante el periodo de mezcla invernal (<14ºC) hasta las elevadas

temperaturas superficiales durante la estratificación estival (>18ºC) en el

océano adyacente a las costas gallegas. También se puede observar como

en Julio, pero sobre todo Agosto y Septiembre, la temperatura superficial

de las aguas costeras es claramente más fría que las del océano adyacente,

debido al fenómeno de afloramiento. Sin embargo, el ciclo estacional

explica <15% de la variabilidad total del viento costero en el afloramiento

de la Península Ibérica, mientras que >70% de dicha variabilidad se

concentra en periodos <30 días (Álvarez–Salgado et al. 2002). Es

especialmente reseñable la variabilidad observada en el comienzo y fin de

la época de afloramiento: mientras que la primera puede ocurrir dentro

del dilatado margen de tres meses, la segunda suele ocurrir en el

relativamente estrecho periodo de un mes (Álvarez–Salgado et al. 2002;

2003).

Page 35: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

17

Figura 6. Sucesión de medias mensuales (Enero a Diciembre) de la temperatura superficial de las costas gallegas y el océano adyacente para el año 2002. Colores “cálidos” en la imagen de satélite (amarillo, naranja y rojo) indican temperaturas progresivamente mayores. Por el contrario, colores “fríos” (verde, azul y morado) indican temperaturas progresivamente menores. Imágenes cedidas por S. Groom, Remote Sensing Group, Plymouth Marine Laboratory (Reino Unido).

Page 36: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

18

2. Geomorfología de las Rías Baixas

La Ría de Vigo es la más meridional de las cuatro Rías Baixas, formadas

por hundimiento de los valles fluviales del Oitabén–Verdugo (Vigo),

Lérez (Pontevedra), Ulla y Umia (Arousa) y Tambre (Muros) durante la

Transgresión Flandriense hace 18000 años. Vidal Romaní (1984) agrupa a

las rías gallegas en cuatro clases (Figura 7):

1. Rías en embudo. El factor fluvial es el principal agente formador de

la ría. Realmente la ría se corresponde con la parte inferior del río

inundada por el mar: rías de Barqueiro, Foz, Ferrol y Ares.

2. Rías tectónicas. En su origen predomina el factor tectónico, su

morfología no puede atribuirse únicamente a la acción erosiva del río: rías

de Vigo, Pontevedra y Muros.

3. Rías de cuenca de alteración terciaria inundadas. Son las rías

verdaderas ya que durante las fases marinas regresivas sufrieron una

importante erosión fluvial.

4. Rías mixtas. Se encuadran aquí las rías que presentan características

de los otros tres grupos citados anteriormente. Se incluyen en este grupo

las rías de Arousa Ortigueira, Cedeira y Viveiro (que tienen caracteres de

1) y 3), mientras que Ribadeo, Laxe, Camariñas y Betanzos tienen

caracteres de 1) y 2).

Las Rías Baixas son bahías en forma de V, que se van haciendo más

anchas y profundas según nos desplazamos hacia la plataforma, donde

llegan a tener hasta 17 Km de ancho y 60 m de profundidad. Este tipo de

morfología favorece el intercambio de agua con la plataforma continental

adyacente. En la Tabla 1 se resumen las características morfológicas más

relevantes de las cuatro Rías Baixas. La Ría de Vigo tiene volumen y

profundidad intermedios, es la más larga y estrecha y la menos profunda.

Page 37: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

19

Vigo

Pontevedra

Arousa

Muros-Noia

Rías Centra

is

Rías Altas

Rías

Baixas

Corcubión

Betanzos

Camariñas

Corme-Laxe

A Coruña AresFerrol

O Barqueiro Foz

RibadeoOrtigueria

Viveiro

Figura 7. Situación de las principales rías gallegas y su clasificación tectónica

según Torre Enciso (1958).

El intercambio de agua entre las rías y la plataforma está parcialmente

bloqueado por la presencia de las Islas Cíes en Vigo, Islas de Ons en

Pontevedra e Isla de Sálvora en Arousa, que conforman el Parque

Nacional Marítimo–Terrestre de las Islas Atlánticas. La boca Sur de estas

tres rías es mucho más ancha y profunda que la boca Norte, de forma que

la mayor parte del intercambio con la plataforma se efectúa a través de

ella. La orografía de las Rías Baixas (Figura 8), con montes relativamente

elevados (mayores de 400 m) flanqueándolas, las protegen del efecto

directo de los vientos, que resultan mucho más débiles que en la

plataforma continental, y orientados a lo largo del eje principal de las rías

(Chase 1975).

Page 38: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

20

Tabla 1. Resumen de la geometría básica de las cuatro Rías Baixas, aportes continentales y tiempo de renovación (media anual) para fines comparativos (Álvarez.- Salgado y Fraga en prensa).

La Ría de Vigo se localiza geográficamente entre los 42º 09’ y los 42º 21’

de latitud Norte y entre los 8º 36’ y 8º 56’ de longitud Oeste (Figura 9).

Desde el punto de vista geomorfológico e hidrográfico, puede dividirse en

tres zonas bien diferenciadas:

1) Zona interna. Comprende la bahía de San Simón, cuenca de

sedimentación de 15 km2 de superficie y 3 metros de profundidad media

donde se descargan la mayor parte de los aportes continentales a la Ría de

Vigo, principalmente el Río Oitabén–Verdugo, parcialmente controlado

por la Presa de Eiras, en Fornelos de Montes. En marea baja, queda al

descubierto el 60% de su superficie.

2) Zona media. Va desde el estrecho de Rande hasta la línea que une

punta Borneira con cabo de Mar. En esta zona se encuentran las

instalaciones de la Autoridad Portuaria de Vigo. Se trata de un canal de

unos 15 kilómetros de longitud, en forma de V, cuya profundidad máxima

aumenta gradualmente desde los 20 metros en Rande a los 40 metros en

Borneira.

Muros Arousa Pontevedra Vigo

Volumen (Km3) 2,74 4,34 3,45 3,12

Superficie (Km2) 125 230 145 176

Profundidad media (m) 22 19 24 18

Longitud (Km) 18 25 23 33

Anchura en boca (Km) 17,2 8,8 5,7 11,6

Profundidad máxima (m) 45 60 40 55

Caudal fluvial medio (m3 s–1) 54 74 15 26

Tiempo de renovación medio (d) 7 18 26 13

Page 39: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

21

Figura 8. Orografía y batimetría de la Ría de Vigo. Tomada de Souto (2000)

3) Zona externa. Va desde la línea Borneira–Cabo de Mar hasta las Islas

Cies, donde la ría se ensancha y gana profundidad. Se comunica con el

Océano Atlántico por la Boca Sur, la Boca Norte y el Freu da Porta

(Figuras 8 y 9). La Boca Norte se sitúa entre la Punta Robaleira y la Isla de

Monte Agudo (o Isla Norte), perteneciente a las Islas Cíes, con una

longitud de aproximadamente 2,5 Km y una profundidad máxima de 25

m. La Boca Sur se sitúa entre Cabo Vicos en la Isla de San Martín (es la isla

Sur de las Cíes) y Monteferro, su longitud es prácticamente el doble de la

Boca Norte y su profundidad máxima ronda los 60 m. El Freu da Porta es

un pequeño y estrecho canal de 7 m de profundidad que separa la Isla de

San Martín de la Isla del Faro (Figura 9).

Page 40: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

22

9.0º 8.9º 8.8º 8.7ºLongitud (W)

42.1º

42.2º

42.3º

Latit

ud (N

)

Rande

C. MarISLAS CÍES

Pta.Borneira

Monteferro

Cabo Estai

La Guía

VIGO

Pta.Robaleira

Freu da Porta

Islas Estelas

I.Toralla

I. de Monte Agudo

I. de S.MartínRío Lagares

RíoRedondela

R.O

itabé

n

MOAÑA

CANGAS

PenínsulaIbérica

E.M. Bouzas

E.M. Peinador

Mareógrafo

Figura 9. Mapa de la Ría de Vigo. Se muestran los principales accidentes geográficos así comos las posiciones de las Estaciones Meteorológicas (E. M.) de Bouzas y Peinador.

3. Meteorología de las Rías Baixas

En la Figura 9 se muestran la posición de la estación meteorológica que

el Instituto Español de Meteorología tiene en la Ría de Vigo, situada en la

terraza del Instituto de Investigacións Mariñas (CSIC) en Bouzas. De esta

estación se han extraído los datos que se presentan en las Figuras 10 y 11.

Temperatura del aire y régimen de precipitaciones de la Ría de Vigo

En la Figura 10 se presentan los valores medios mensuales, así como

los límites dentro de los que se encuentran el 50% y el 80% de las

observaciones de temperatura del aire (en ºC) y precipitación (en mm d-1)

registrados en la estación meteorológica de Bouzas durante el periodo

1993–2003. Los valores medios anuales de temperatura del aire oscilan

entre 9,6ºC en Enero y 18,7ºC en Agosto. Los rangos de variación de

temperatura en el que se encuentran el 80% de las observaciones durante

Page 41: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

23

esos 10 años son muy amplios: 5,4 – 13,4 ºC en Enero y 15,1 – 22,7 ºC en

Julio.

Los valores medios anuales de precipitación oscilan entre 1,1 mm/d en

Junio y 6,4 mm/d en Noviembre. Igualmente, los rangos de variación de

precipitación en el que se encuentran el 80% de las observaciones durante

esos 10 años son muy amplios: 0 – 22,4 mm d-1 en Noviembre y 0 – 20,2

mm d-1 en Marzo.

En la época estival la temperatura del aire, y por tanto la irradiancia

adquiere mayor importancia relativa como motor de la circulación

gravitacional, tal como muestran los trabajos de Álvarez-Salgado et al.

(2000) y Gilcoto et al. (2001). Debido al calentamiento superficial la

irradiancia genera un gradiente vertical de temperatura en la columna de

agua convirtiéndose en la principal fuente de estratificación durante los

meses de verano.

En la Figura 11 se muestra la intensidad y dirección más probables del

viento local registrado en la estación meteorológica de Bouzas en el

periodo 1993–2003. Los colores más cálidos indican aumento progresivo

de la probabilidad. Por el contrario, los colores fríos indican descenso

progresivo de la probabilidad. Se observa que los vientos más probables

suelen ser de intensidad menor a 4 m s-1 soplando en la dirección del eje

de la ría (vientos del NE, en sentido saliente, o del SW, en sentido

entrante). La reducción de la intensidad del viento en la ría respecto a su

intensidad en plataforma, así como su orientación respecto al eje principal

de las rías fue inicialmente observada por Chase (1975).

Page 42: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

24

E F M A M J J A S O N D

t aire

(ºC)

0

5

10

15

20

25

30

1 14 27 40 53

T air

e (ºC

)

8

10

12

14

16

18

20

E F M A M J J A S O N D

prec

ipita

ción

(mm

d - 1

)

0

20

40

60

80

100

Figura 10. Valores mensuales de la temperatura del aire y la precipitación en la Estación Meteorológica de Bouzas obtenidos a partir de una serie diaria de datos de 1993 a 2003. En las gráficas (a) y (b) se muestran la media (línea de puntos), mediana (línea continua), el rango en el que se encuentran el 50% de las observaciones (caja) y el rango en el que se encuentran el 80% de las observaciones (bigotes). En la grafica c se muestra el valor medio semanal real (puntos) y suavizado (líneas) en el periodo 1993–2003 para la temperatura del aire.

Page 43: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

25

También son importantes los vientos locales en dirección transversal al

eje de la Ría de Vigo, es decir, en sentido Vigo–Cangas (vientos de

componente Sur) o Cangas–Vigo (vientos de componente Norte).

Aparte hay que considerar el régimen de brisas, tierra–mar durante la

noche y mar–tierra durante el día (Villarino et al. 1995).

8 m/s

7

6

5

4

3

2

1

Norte

Oeste Este

Sur10-3 %

10-2 %

10-1 %

100 %

Figura 11. Rosa de los vientos locales en la Ría de Vigo, a partir del registro horario de la Estación Meteorológica de Bouzas en el periodo 1993–2003. La escala de colores indica la probabilidad de cada intensidad (mayor intensidad a mayor radio de los círculos concéntricos) y dirección y sentido (N, S, E y O) dados.

Page 44: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

26

Los aportes continentales a la Ría de Vigo son el resultado de la

combinación de un flujo regulado por la presa de Eiras, en el curso alto

del Río Oitabén (Figura 9), que suministra agua potable a la Ría de Vigo, y

un flujo natural de agua, debido fundamentalmente al Río Verdugo, que

confluye con el Oitabén antes de desembocar en la Bahía de San Simón y,

en menor medida los ríos Redondela y Ullo (Gago et al. 2005).

En la Figura 12 se representan los valores medios mensuales, así como

los límites dentro de los que se encuentran el 50% y el 80% de las

observaciones de aportes continentales a la zona interna y media de la Ría

de Vigo (en m3 s-1). Presenta un ciclo estacional muy marcado con un

máximo invernal y un mínimo estival. Los valores medios anuales de

aportes continentales oscilan entre 9 m3 s-1 en Agosto y 42 m3 s-1 en

Noviembre. Los rangos de variación en el que se encuentran el 80% de las

observaciones realizadas son muy amplios: 6 – 95 m3 s-1 en Diciembre.

Page 45: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

27

E F M A M J J A S O N D

caud

al río

(m3 ·s

-1)

0

59

118

176

235

1 14 27 40 53

caud

al río

(m3 s-1

)

0

12

24

35

47

59

Figura 12. Valores mensuales de los aportes continentales a la zona interna y media de la Ría de Vigo para el periodo 1987–2000. En la grafica (a) se muestran la media (línea de puntos), mediana (línea continua), el rango en el que se encuentran el 50% de las observaciones (caja) y el rango en el que se encuentran el 80% de las observaciones (bigotes). En la gráfica b se muestra el valor medio semanal real (puntos) y suavizado (líneas) en el periodo en el periodo 1987–2000.

a

Page 46: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

28

4. Circulación de las Rías Baixas

La circulación de la Ría de Vigo consta de dos componentes: mareal

(generada por las mareas) y residual (independiente de la marea, función

de los aportes continentales y los vientos locales y de plataforma).

Los movimientos periódicos asociados al fenómeno mareal tienen su

origen en la elevación de la superficie libre en la frontera con la

plataforma (Pugh 1996). El cambio de altura del nivel del mar se traduce

en una oscilación de las corrientes que en zonas confinadas como las rías

pueden llegar a ser de gran intensidad y provocar una fuerte mezcla

vertical (Míguez 2003). La mayor parte de la energía implicada en la

dinámica mareal se concentra en períodos semidiurnos (dos bajamares y

dos pleamares a lo largo del día).

Los armónicos de mayor importancia en el rango semidiurno en

términos de amplitud son: lunar principal, M2, seguida de solar principal,

S2, que la modula con un período de medio mes lunar originando mareas

vivas y muertas. La componente N2 modula las mareas vivas y muertas

haciéndolas más o menos intensas, y por lo tanto es la responsable de las

desigualdades semimensuales tanto entre mareas vivas como muertas

(Míguez 2003). Así, la amplitud de las dos mareas que se producen cada

día pasa, en unos quince días, de los 2 metros en una situación de marea

viva a los 50 cm en una de mareas muertas (Figura 13).

Para el análisis de la corriente de marea los registros de corriente total

se someten a un proceso de filtrado con el fin de eliminar la influencia de

los forzamientos de baja frecuencia. En la Ría de Vigo, la importancia de

las componentes astronómicas de baja frecuencia es muy inferior a la de

las componentes diurnas y semidiurnas (Míguez 2003). Por lo tanto, los

registros de corrientes se someten a un filtro pasa baja A242A25 (Godin

Page 47: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

29

1972) y el resultado de este filtrado se resta a la serie de corriente original.

Se obtiene una nueva serie en la que sólo quedan registrados aquellos

procesos con período inferior a 30 horas, seleccionándose de esta forma las

frecuencias diurnas y semidiurnas.

La Figura 14 muestra el análisis vectorial armónico de las corrientes de

marea a dos profundidades (13 m y 26 m) para los armónicos de

frecuencia semidiurna más energéticos (M2 y S2). Para cada frecuencia el

vector velocidad se parametriza por sus semiejes mayor y menor, el

ángulo de inclinación del semieje mayor respecto al eje X (Este) y la fase

referida a Greenwich. El hecho de que el semieje mayor del armónico más

energético (M2) sea mucho mayor que el semieje menor implica que la

corriente de entrada (llenante) y la corriente de salida (vaciante) siguen

prácticamente la misma dirección. Se observa también que las elipses

están orientadas siguiendo la geometría de la Ría de Vigo.

Para el análisis de la corriente residual los registros de corriente total se

someten a un proceso de filtrado con el filtro pasa baja A242A25 (Godin

1972) de modo que se eliminan los procesos que transcurren con

frecuencias superiores a 1 día.

Page 48: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

30

Figura 13. Variación del nivel del mar en la Ría de Vigo en 2001 medida cada hora en el puerto de Vigo (datos tomados por Puertos del Estado, www.puertos.es). En las gráficas se presentan la variación del nivel del mar descompuesto en una serie meteorológica (residual, línea verde) y en una serie astronómica (mareal, línea violeta), la suma de ambas es el nivel del mar medido. La gráfica (a) representa el periodo del 1 de Junio al 8 de Agosto del año 2001; la gráfica (b) representa una ampliación del período del 15-Junio a 12-Julio, periodo correspondiente al fondeo de un correntímetro ADCP DCM12 en una estación situada entre C. de Mar y Punta Borneira en la Ría de Vigo (Figura 9).

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

1-jun 8-jun 15-jun 22-jun 29-jun 6-jul 13-jul 20-jul 27-jul 3-ago

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

15-jun 22-jun 29-jun 6-jul 13-julAltu

ra d

e m

area

(m)

Page 49: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

31

Figura 14. Elipses de la velocidad de corriente de marea de los armónicos semidiurnos M2 y S2, correspondientes a las corrientes de marea registradas por el correntímetro doppler DCM12 a 13m y 26m de profundidad durante el periodo 15 Junio-12 Julio de 2001. Se muestrean los parámetros básicos de las elipses: M (semieje mayor); m (semieje menor); I (inclinación); θ (fase).

En la Figura 15 se observa cómo el carácter periódico de la marea

implica que tenga poca importancia en el transporte de agua a escalas

superiores a las 6 horas (Figueiras et al. 1994; Montero 1999).

Como el objetivo de esta memoria se concentra en los procesos de

intercambio de materia y energía entre la ría y la plataforma adyacente,

M2 S2

M(cm/s) 10.2±1.8 2.4±1.7

m(cm/s) 0.3±1.7 0.2±1.6

I(º) 28±11 23±41

Θ(º) 24±10 51±52

M2 S2

M(cm/s) 11.5±2.1 2.5±1.6

m(cm/s) -0.3±1.7 -0.2±1.6

I(º) 32±8 36±39

Θ(º) 18±11 56±47

Page 50: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

32

los fenómenos físicos en los que nos centraremos son aquellos que tienen

la energía asociada a la baja frecuencia, y que dan lugar a la circulación

residual, que es la que realmente influye en el trasporte de agua a largo

plazo. Míguez (2003) ha profundizado en la corriente mareal. De todos

modos no se descartan futuros trabajos cuyo objeto de estudio sea la

corriente de marea obtenida a partir de las series de corrientes

correspondientes a los fondeos de correntímetros efectuados en el

transcurso del proyecto REMODA.

10 -5 0 5

-25

-20

-15

-10

-5

0

5

-10 -5 0 5

-25

-20

-15

-10

-5

0

5

-10 -5 0 5

-25

-20

-15

-10

-5

0

5

Km Km Km

Km Km Km

Figura 15. Trayectoria virtual de la circulación total (a), residual (b) y mareal (c) en superficie registrada por un correntímetro doppler en la Ría de Vigo. Se observa que la marea tiene poca importancia en el transporte de agua a largo plazo debido a su carácter periódico.

5. Hidrodinámica y biogeoquímica de las Rías Baixas

El análisis cuantitativo del origen y destino del material biogénico

que se produce en las rías exige la aplicación de 1) balances de materia

basados en medidas in situ de corrientes residuales, variables

termohalinas (salinidad, temperatura, densidad) y químicas (oxígeno

a b c

Page 51: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

33

disuelto, carbono inorgánico, alcalinidad, sales nutrientes, materia

orgánica disuelta y particulada, etc.) o tasas de sedimentación con 2)

medidas in vitro para determinar la producción primaria, respiración,

exudación de materia orgánica disuelta, herbivoría, etc. Afortunadamente,

desde mediados de la década de 1970 se han llevado a cabo estudios en las

rías aplicando las dos aproximaciones, si bien no siempre de forma

simultánea, de forma que las rías gallegas son sistemas de afloramiento

costero bien conocidos.

5.1. Conocimiento hidrodinámico previo al proyecto REMODA

La primera aproximación a la hidrodinámica de las rías gallegas se

basó en los balances de cajas estacionarios aplicados inicialmente a la

salinidad. Posteriormente, Otto (1975) los aplicó a la temperatura. Estos

modelos estacionarios como el de Otto (1975) o el de Prego y Fraga (1992)

asumían que ∆S/∆t = 0 y ∆T/∆t = 0. Posteriormente Rosón et al. (1997)

planteó las ecuaciones de balances de agua, sal y temperatura

considerando el carácter transitorio de las rías. En base a la Figura 16 se

pueden escribir los balances de agua, sal y temperatura para una ría

(Rosón et al. 1997, Álvarez-Salgado et al. 2000):

EPQQQ RBS −++=

SSBB SQSQtSV ⋅−⋅=

∆∆

HT)EPQ(TQTQtTV RRSSBB +⋅+++⋅−⋅=

∆∆

(1)

(2)

(3)

donde QR es el caudal del río (en m3 s-1), QS e QB son los caudales

residuales de superficie y fondo en la boca de la ría (en m3 s-1), V es el

volumen de la ría (en m3), SS y SB son las salinidades de superficie y fondo

en la boca da ría, TR es la temperatura del río, TS y TB son las temperaturas

Page 52: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

34

(en ºC) de superficie y fondo en la boca de la ría. H es el intercambio de

calor con la atmósfera (en ºC m3 s-1). Por último, ∆S /∆t en s–1 y ∆T/∆t en

ºC s-1 son los cambios de salinidad y temperatura en el tiempo.

P-E CQ

QRMZQE

QZQS

QB

capa superficial

capa de fondo

P-E CQ

QRMZQE

QZQS

QB

capa superficial

capa de fondo

Los modelos de cajas en su aproximación no estacionaria aumentaron

el conocimiento sobre la hidrodinámica de las rías gallegas y su relación

con los forzamientos externos. Así, los modelos de cajas que tenían como

fuente datos de salinidad y temperatura obtenidos con periodicidad de ½

semana han sido muy útiles para estudiar el acoplamiento entre el viento

que sopla en plataforma y la circulación residual de la Ría de Arousa

(Rosón et al. 1997), de la Ría de Vigo (Álvarez-Salgado et al. 2000) y

posteriormente al proyecto REMODA, de la Ría de Pontevedra (Pardo et

al. 2001).

En las Figuras 17 y 18 se muestran los caudales medios de agua

continental y oceánica que circulaban por la Ría de Arousa entre Mayo y

Octubre de 1989 a partir de una serie de 46 muestreos consecutivos cada

3–4 días a los que se aplicaron los balances de agua, sal y temperatura

Figura 16. Representación esquemática de una ría. Se presenta una situación de circulación residual positiva. QS y QB, caudales horizontales de entrada (salida) en el nivel inferior (superior). QZ, caudal neto de ascenso hacia la superficie. MZ, intercambio por mezcla turbulenta entre capas. H, entrada neta de calor desde la atmósfera hacia la superficie; P, precipitación; E, evaporación; y QR, caudal del río en la capa superficial

Page 53: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

35

(Álvarez–Salgado et al. 1996a; 1996b; Rosón et al. 1999; Pérez et al. 2000b).

En la figura 17 se muestran los caudales medios durante en período de

afloramiento. El 99,4% del agua que circula por la ría viene de la

plataforma continental adyacente y las aguas continentales sólo

representan el 0,6%. La convección y la difusión turbulenta vertical son de

magnitud comparable.

P-E

MZQZ

QS

QB

QR

5069

5069

5097

0.5

285205

P-E

MZQZ

QS

QB

QR

5069

5069

5097

0.5

285205

Figura 17. Flujos en la Ría de Arousa del 08/06 al 15/10/1989. Caudales netos (en m3 s-1). Adaptado de Pérez et al. (2000b).

En respuesta a los vientos de componente Sur en la plataforma, la

circulación residual de la Ría de Arousa se invierte, con entrada de agua

superficial de la plataforma y del río y salida de agua del fondo de la ría

hacia la plataforma (Figura 18). En condiciones de hundimiento se forma

un frente de convergencia en el interior de la ría. En este caso el 99,9% del

agua que circula por la ría es agua superficial de la plataforma (Álvarez–

Salgado et al. 1996a; Rosón et al. 1999).

Page 54: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

36

QR831812832

QB

QS

P-E

MZQZ

12632

12618

0.8

14QR

831812832

QB

QS

P-E

MZQZ

12632

12618

0.8

14

Figura 18. Flujos y balances medios de nitrógeno en la Ría de Arousa del 15/10 al 30/10/1989. Caudales netos (en m3 s-1). Adaptado de Pérez et al. (2000b).

Los modelos de cajas no estacionarios aplicados a las rías encajan

perfectamente en la teoría de los métodos inversos, lo que llevó mejorar la

solución de los mismos mediante la aplicación del método OERFIM

(Optimum Esturarine Residual Fluxes with a multiparameter Inverse

Method) en la Ría de Vigo (Gilcoto et al. 2001). Este método, aplicado

posteriormente a la solicitud del proyecto REMODA, se diseñó para

aplicarlo a una geometría bidimensional y en una única caja en forma de

cuña dividida verticalmente en dos semicajas, de manera que se adaptase

a la geometría de las rías gallegas y a la tradicional circulación residual en

doble capa asociada a los estuarios parcialmente mezclados.

Todos estos estudios concluían que los vientos costeros eran los

principales responsables de los cambios hidrográficos de las rías, y estos

cambios eran compatibles con la suposición de circulación residual en

doble capa previa a la ejecución de los mismos.

En las rías gallegas, la información hidrodinámica obtenida a partir de

medidas directas de corrientes ha sido escasa hasta la fecha. Los primeros

estudios fueron llevados a cabo mediante el fondeo de correntímetros

Page 55: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

37

mecánicos por Castillejo et al. (1982) en la Ría de Arousa, y Figueiras et al.

(1994) y Villarino et al. (1995) en la Ría Vigo. De todos modos las medidas

no cubrían verticalmente toda la columna de agua como para confirmar la

circulación residual en doble capa, es decir los datos de corrientes eran

muy dispersos y nunca se efectuaron medidas continuas en un punto

durante periodos suficientemente largos y en situaciones oceanográficas

distintas (en las distintas fases del ciclo de estratificación/homogenización

de las rías).

La Figura 19 muestra datos en continuo de temperatura del agua y de

la corriente residual superficial recogidos en la estación situada en el

centro de la sección trasversal Cabo de Mar – Punta Borneira de la Ría de

Vigo a lo largo del mes de Septiembre de 1990. La gráfica superior (Figura

19a) muestra la evolución temporal de la componente Norte del viento

que sopla en la plataforma adyacente. Se observa una clara evolución

desde 1) una situación de calma el 14 de Septiembre a 2) viento Norte

(afloramiento) el 17 de Septiembre, seguido de 3) una relajación que se

extiende hasta el 24 de Septiembre y, finalmente, 4) un intenso episodio de

viento Sur (hundimiento), que alcanza su cenit el 27 de Septiembre.

La gráfica central (Figura 19b) muestra la respuesta de la columna de

agua al viento costero. La evolución temporal de las isotermas de 14–15ºC

es paralela a la del viento Norte desde el 12 al 24 de Septiembre. Durante

el episodio de hundimiento del 24 al 27 de Septiembre el agua de 14–15ºC

se retira de la columna de agua, pasando a ser ocupado su lugar por agua

de más de 17ºC en los primeros 20m. La gráfica inferior (Figura 19c)

muestra la trayectoria virtual de la corriente superficial medida por el

correntímetro en la anterior estación hidrográfica. Se ha aplicado un filtro

Page 56: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

38

Figura 19. Evolución temporal de la componente Norte del viento en plataforma (a), temperatura de los 25 primeros metros de la columna de agua (b), y trayectoria virtual de la circulación residual en superficie registrada por un correntímetro mecánico RCM7 a 4m de profundidad (c) en el segmento central de la Ría de Vigo. Tomada de Gilcoto et al. (2001).

Page 57: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

39

de marea a los datos del correntímetro, de manera que la trayectoria

representada se corresponde con la circulación residual, es decir, aquella

que queda una vez se elimina el efecto de la marea. Gilcoto et al. (2001),

posteriormente a la solicitud del proyecto REMODA, mostraron cómo la

circulación residual medida con datos reales de corriente en la Ría de

Vigo, y por lo tanto la distribución de las variables termohalinas, están

muy afectadas por los procesos de afloramiento y hundimiento

desarrollados en la plataforma.

En la Tesis Doctoral de Míguez (2003) se avanzó en el conocimiento de

la hidrodinámica de las rías mediante el análisis exhaustivo de las series

de corriente obtenidas del fondeo de correntímetros mecánicos y doppler

en las Rías de Pontevedra y de Vigo dentro del proyecto Ordenación

Integral del espacio marítimo-terrestre de Galicia, subprograma: Dinámica

y variabilidad termohalina (FEUGA, 1996-1999). La presente Tesis,

enmarcada en el Proyecto REMODA, se centra en el conocimiento

dinámico de la Ría de Vigo en su zona central cubriendo situaciones

oceanográficas diferentes, y da un paso más al intentar caracterizar las

diferencias laterales en las componentes longitudinales y transversales de

la circulación residual de la Ría de Vigo a partir del análisis, efectuado por

primera vez, de datos de ADCP (Acoustic Doppler Current Profiler)

obtenidos a bordo de del B/O Mytilus.

Es necesario destacar la contribución que han realizado los modelos

numéricos al conocimiento hidrodinámico de las rías gallegas. Las

primeras aportaciones fueron las de Pascual y Calpena (1985) y Pascual

(1986) que caracterizaron la circulación mareal y forzada por el viento

local en la Ría de Arousa. El cálculo de los modelos numéricos mejoró

notablemente en la década de 1990 debido a una mayor resolución

Page 58: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

40

espacial, pero tenían carencias al no considerar el viento en plataforma

como forzamiento. Entre los trabajos con aplicación de modelos

numéricos están los de Bruno et al. (1998) que proporcionó un método

para la predicción de corriente de marea a las Rías de Pontevedra y de

Vigo, Montero et al. (1999), que estudió la circulación residual en la Ría de

Vigo mediante un modelo 3D baroclínico y Gómez-Gesteira (1999), que

aplicó un modelo 2D para el estudio de la dispersión de contaminantes en

las Rías de A Coruña y Vigo.

Posteriormente a la solicitud del proyecto REMODA, el conocimiento

de la circulación residual de la Ría de Vigo inducida por el viento costero

mejoró notablemente con la aplicación de los modelos GHER

(GeoHydrodynamics and Environment Research) y HAMSOM

(HAMburg Shelf Ocean Model) por Torres-López et al. (2001) y Souto et

al. (2001), respectivamente. Ambos plantearon que existía una circulación

residual en dos capas en la parte central de la ría y una circulación

residual tridimensional en la parte más externa de la ría, que se debía a la

interacción del viento de plataforma con la particular topografía de la

parte externa de la ría (dos bocas de entrada separadas por las Islas Cíes).

Posteriormente, Souto et al. (2003) confirmó dicha circulación comparando

datos experimentales y simulados. Esta circulación tridimensional ha sido

recientemente encontrada por Gilcoto et al. (2006) mediante la aplicación

de un modelo de caja tridimensional. Gilcoto et al. (2006) comprobó que la

circulación tridimensional en la parte externa afecta a la parte interna de la

ría dando lugar a variaciones transversales en el flujo, tal como se muestra

en el capítulo 3 de esta memoria, donde se describen estas variaciones a

partir de los registros de corrientes del ADCP instalado a bordo del B/O

Mytilus.

Page 59: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

41

5.2. Conocimiento biogeoquímico previo al proyecto REMODA

La circulación residual i) transporta por convección y difusión

turbulenta las sales nutrientes que fertilizan las rías ii) determina el

tiempo que esas sales nutrientes están a disposición del plancton

(bacterias, fitoplancton, protozoos y metazoos) que habita la ría y iii)

condiciona el destino del material producido por el fitoplancton

(respiración, disolución, exportación horizontal, sedimentación,

preservación en los fondos da ría, transferencia a organismos superiores,

etc.).

Las rías pueden asimilarse a grandes reactores en los que los reactivos

(sales nutrientes) se transforman en productos (material biogénico) y

viceversa, debido a la intervención de todos los organismos que viven en

el reactor, desde las bacterias, pasando por el pico–, nano– y

microplancton, los protozoos (ciliados y flagelados) hasta las macroalgas,

el zooplancton y los herbívoros y carnívoros superiores (moluscos,

crustáceos y peces).

5.2.1. Impacto de la hidrografía y la dinámica en los ciclos biológicos

y geoquímicos de las Rías Baixas

Los procesos biológicos y geoquímicos que ocurren en las rías pueden

agruparse en dos grandes categorías: 1) procesos de biosíntesis, en los que

se produce materia biogénica a partir de las sales nutrientes y 2) procesos

de mineralización, en los que se degrada la materia biogénica para

producir nuevamente sales nutrientes.

El proceso de biosíntesis más relevante es la fotosíntesis, ejecutada por

las macroalgas (10%) y, sobre todo, por el fitoplancton (90%; Varela et al.

1984), en la capa superficial de las rías, donde la intensidad de radiación

Page 60: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

42

útil para la fotosíntesis (longitudes de onda entre 400 y 700 nm) es

superior al 1% de la intensidad en la interfase aire–agua (Tett y Edwards

1984). Durante la biosíntesis se produce fundamentalmente materia

orgánica y estructuras silíceas (SiO2) y calcáreas (CaCO3).

Aunque la composición del fitoplancton marino es muy variable lo

cierto es que si se consideran escalas espaciales (decenas de kilómetros) y

temporales (meses) suficientemente amplias, dicha composición no difiere

mucho de C106H171O44N16P. Esta composición media equivale a los % de

las principales biomoléculas que se muestran en la Tabla 2 (Anderson

1995; Fraga et al. 1998; Fraga 2001).

Usando esta composición como producto de la fotosíntesis, la reacción

completa de los procesos de biosíntesis en nuestro reactor biogeoquímico

será:

→β+α+++ +−−− 244

2433 CaSiOHHPONO16HCO230 (4)

OH32CO)124(O148)CaCO()SiO)(PNOHC( 2232321644171106 +β−++→ −

βα

A diferencia de los procesos de biosíntesis, los procesos de

mineralización, que incluyen la respiración de la materia orgánica y la

disolución de las estructuras de SiO2 y CaCO3 biogénicos, ocurren en toda

la columna de agua y en los sedimentos. Los productos finales de la

mineralización son las sales nutrientes originales (HCO3-, NO3-, HPO42- y

H4SiO4), cerrándose el ciclo de estos elementos en el medio marino. En la

respiración de la materia orgánica toman parte todos los organismos que

habitan la ría.

Page 61: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

43

Chemical formula

% (w/w)

Compuestos de fósforo C45H76O31N12P5 12

Proteinas C139H217O45N39S 45,1

Pigmentos fotosintéticos C46H52O5N4Mg 2

Lípidos C53H89O6 16,5 Hidratos

de Carbono C6H10O5 24,4

Average composition C106H171O44N16P 100

Tabla 2. Composición media de los principales grupos de biomoléculas que constituyen la materia orgánica del fitoplancton marino. Igualmente se indica la proporción (en peso) en que participa cada grupo a la composición media del fitoplancton. Tomada de Fraga (2001).

En lo referente a la disolución de estructuras silíceas y calcáreas se trata

de un mero proceso fisicoquímico, que depende básicamente del pH, la

temperatura del agua y la presión hidrostática. Mientras que en el rango

de pH del agua de mar, la acidez no tiene efecto en la solubilidad de la

sílice, la temperatura favorece su disolución al aumentar su solubilidad y

la velocidad específica de disolución (Spencer 1983). Desde un punto de

vista termodinámico, la disolución de CaCO3 no está favorecida donde

haya sobresaturación. El descenso de pH asociado a la mineralización de

materia orgánica reduce la [CO32-], de modo que producto [CO32-]·[Ca2+]

llega a estar por debajo de la saturación a pH <7,8 para el aragonito y <7,6

para la calcita. Sin embargo, la disolución masiva de CaCO3 sólo comienza

cuando [CO32-] es <70% de la saturación (Broecker y Peng 1982).

El fitoplancton y los microheterótrofos (bacterias heterotróficas,

ciliados y flagelados) constituyen la llamada “red trófica microbiana”

Page 62: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

44

(Figura 20; Azam et al. 1983) en la que el material biogénico producido por

el fitoplancton queda atrapado sin ser de utilidad para la explotación de

los recursos vivos de las rías. Los microheterótrofos desempeñan el papel

más relevante en la mineralización de la materia orgánica (Harrison 1980).

El fitoplancton y los microheterótrofos son, en su conjunto, una fuente de

alimento para la “cadena trófica de los metazoos” (Sherr y Sherr 1988),

que comienza en el zooplancton herbívoro y acaba en formas de vida

explotables tales como moluscos, crustáceos y peces. Por consiguiente, la

dominancia de la “red trófica microbiana” o de la “cadena trófica clásica”

tiene una gran relevancia para la explotación de los ecosistemas marinos.

La Figura 21 muestra la evolución estacional de la concentración de

clorofila, indicador de la biomasa de fitoplancton, en el segmento central

de la Ría de Vigo. Como en todo ecosistema marino de latitudes

templadas, la clorofila superficial se caracteriza por valores mínimos en

invierno y máximos en primavera y en otoño. Sin embargo, las

concentraciones de clorofila superficial en verano son mayores de lo

esperado, debido precisamente a la entrada intermitente de sales

nutrientes a consecuencia de los episodios de afloramiento. Las

proliferaciones de primavera y otoño coinciden con los períodos de

transición de la época de afloramiento a la de hundimiento y viceversa

(Figura 2).

Page 63: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

45

fito

plan

cton

bucl

em

icro

bian

o

< 5 µm

0.1 - 2 µmbacterias

2 - 30 µmflagelados

8 - 100 µmciliados

MODy

detritus

met

azoo

s

> 5 µm

fito

plan

cton

bucl

em

icro

bian

o

< 5 µm

0.1 - 2 µmbacterias

2 - 30 µmflagelados

8 - 100 µmciliados

MODy

detritus

met

azoo

s

> 5 µm

Figura 20. Representación esquemática de la “red trófica microbiana” y sus conexiones con la “cadena trófica clásica” de los metazoos. Se considera el fitoplancton dividido en dos fracciones <5 µm y >5 µm. Únicamente las células > 5 µm sirven de alimento a los metazoos. MOD es la materia orgánica disuelta que se produce por exudación y lisis celular. Adaptado de Sherr y Sherr (1988).

Estos períodos de transición se caracterizan por una gran inestabilidad

atmosférica, de forma que vientos de componente Norte pueden rolar a

Sur y viceversa en días. Según la dirección del viento dominante, se

favorece la exportación hacia el océano o el confinamiento en las rías

(Figuras 4 y 5) del material biogénico producido por el fitoplancton

durante las proliferaciones primaveral y otoñal (Álvarez–Salgado et al.

2003, 2006; Arístegui et al. 2006)

Page 64: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

46

Chl

-a (m

g m

-3)

02468

10

X F M A M X X A S O N D

El ciclo estacional del fitoplancton (Figura 22), consta de 1) un período

invernal dominado por nanoplancton (más del 60% tiene entre 2 y 20 µm)

y diatomeas bentónicas; 2) un período de proliferación primaveral

dominado por diatomeas (microplanton, >20 µm); 3) un período estival en

el que se suceden, por este orden, diatomeas, dinoflagelados y pequeños

heterótrofos, en función de la intensidad y duración de los ciclos de

afloramiento/relajación/calma, que se repiten cada 10–20 días (Pazos et

al. 1995); y 4) un período de proliferación otoñal dominado por

dinoflagelados migradores (microplancton > 20 µm; Figueiras y Ríos 1993)

Como puede apreciarse en la Figura 21 mientras que la proliferación

primaveral de diatomeas produce un máximo de concentración de

clorofila en el fondo debido a la sedimentación, la proliferación otoñal de

dinoflagelados se mantiene en la superficie. El frente de convergencia que

se forma en el interior de la ría durante el otoño favorece la selección de

los dinoflagelados migradores frente a las diatomeas porque los primeros

tienen capacidad para mantenerse a flote contrarrestando las condiciones

de hundimiento impuestas por los vientos dominantes de componente Sur

(Figueiras et al. 1994; Fermín et al. 1996). Estos dinoflagelados migradores

son la causa de las “mareas rojas” que tanto perjuicio causan a la

explotación de moluscos bivalvos de las rías, que acumulan las biotoxinas

producidas por estos organismos. El picoplancton (<2 µm) es una

Figura 21. Valores medios quincenales de clorofila (en mg/m3) de superficie y fondo en el segmento central de la Ría de Vigo a partir de una serie semisemanal de datos de 1987 a 1996. Se muestra la media (círculos) y la desviación estándar (líneas discontinuas). Círculos blancos, superficie; círculos negros, fondo. Tomada de Nogueira et al. (1997).

Page 65: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

47

componente menor del fitoplancton de las rías, representando un 13±10%

da clorofila total (media anual; Rodríguez et al. 2003).

diatomeasestratificación

prim

aver

a

invierno

verano

otoñ

o

hom

ogen

eiza

ción

reciclado

reno

vaci

ón

dinoflagelados

pequeñosflagelados

ciliados

dinoflageladosDiatomeas

bentónicasnanoplancton

picoplancton

diatomeasestratificación

prim

aver

a

invierno

verano

otoñ

o

hom

ogen

eiza

ción

reciclado

reno

vaci

ón

dinoflagelados

pequeñosflagelados

ciliados

dinoflageladosDiatomeas

bentónicasnanoplancton

picoplancton

Figura 22. Ciclo estacional del plancton en las Rías Galegas. Adaptado de Figueiras y Niell (1987).

En la Figura 23 se muestra la evolución estacional del contenido en

sales nutrientes en superficie y fondo del segmento central de la Ría de

Vigo. Se eligieron las tres formas de nitrógeno inorgánico: nitrito (media

anual 0,32 µmol kg-1 en superficie, 0,54 µmol kg-1 en fondo), amonio

(media anual 1,88 µmol kg-1 en superficie, 2,43 µmol kg-1 en fondo) y

nitrato (media anual 2,99 µmol kg-1 en superficie, 5,76 µmol kg-1 en fondo),

en orden de concentración creciente. Las tres formas presentan una

evolución estacional similar, tanto en superficie como en fondo. En la

superficie, las concentraciones son mínimas en primavera y en verano y

máximas en otoño e invierno, en relación con el ciclo estacional de

consumo de nutrientes por el fitoplancton, típico de los ecosistemas

templados (Figura 21). El consumo es mayor que los aportes oceánicos y

continentales de nutrientes en primavera y en verano y menor en otoño e

invierno. El máximo superficial de amonio ocurre en la segunda quincena

Page 66: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

48

de Octubre, el de nitrito en la segunda de Noviembre y el de nitrato en la

segunda de Diciembre, es decir, con un mes de diferencia. Este

comportamiento se debe a que en los procesos de mineralización, la

primera forma de nitrógeno que se libera es el amonio, que se oxida a

nitrito y el nitrito se oxida a nitrato. La conversión de amonio a nitrito y de

nitrito a nitrato, conocido globalmente como “nitrificación”, lo realizan las

“bacterias nitrificantes” (Wada y Hattori 1991). Esta sucesión de máximos

debido a la secuencia de mineralización: amonio → nitrito → nitrato ha

sido observada en diferentes ambientes marinos (Spencer 1975, Millero y

Sohn 1992)

En el fondo se observa una evolución estacional contraria: las máximas

concentraciones de amonio, nitrito y nitrato ocurren en verano. Las causas

de esta variación son: 1) una mayor entrada de ENACW fría, salada y rica

en nitrato en las rías a consecuencia de la dominancia de los vientos de

componente Norte en la plataforma y 2) la mineralización local en las

aguas del fondo y en los sedimentos del material fitoplanctónico que

sedimenta desde la capa superficial. La mineralización es la causa de los

máximos de amonio y nitrito ya que el ENACW no transporta amonio ni

nitrito (Álvarez-Salgado et al. 2006) y contribuye parcialmente al máximo

de nitrato. Al igual que en la evolución estacional de superficie, en fondo

también existe un desfase temporal en los máximos de amonio, nitrito y

nitrato que, como en superficie, se debe a los dos pasos en que consisten

los procesos de nitrificación. El fosfato y el silicato siguen ciclos

estacionales similares al del nitrato (Nogueira et al. 1997).

Page 67: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

49

amon

io (µ

mol

kg-1

)0

1

2

3

4

5

nitr

ito (µ

mol

kg-1

)

0.00

0.25

0.50

0.75

1.00ni

trat

o (µ

mol

kg-1

)

0

2

4

6

8

10

c

b

a

X F M A M X X A S O N D Figura 23. Valores medios quincenales de amonio (a), nitrito (b) y nitrato (c) de superficie y de fondo en el segmento central de la Ría de Vigo a partir de una serie semisemanal de datos de 1987 a 1996. Se muestran la media (puntos) y la desviación estándar (líneas discontinuas). Puntos blancos, superficie; puntos negros, fondo. Las concentraciones están en µmol kg-1. Tomado de Nogueira et al. (1997).

5.2.2. Balance biogeoquímico de las rías

Los primeros balances biogeoquímicos de las rías gallegas fueron los de

sales nutrientes o materia orgánica llevados a cabo por Otto (1975) y

González et al. (1979) en la Ría de Arousa, y Prego (1993a; 1993b; 1994;

Page 68: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

50

1995; 2002) en la Ría de Vigo. En todos ellos se asumía un comportamiento

estacionario de las rías, es decir, que la concentración de la variable

estudiada (C) no cambiaba en el tiempo ( 0dtdC

= ). Sin embargo, pronto se

descubrió que la aproximación de comportamiento estacionario era

insuficiente para reproducir los ciclos de fertilización/empobrecimiento

en sales nutrientes que se suceden en las rías cada 10–20 días a lo largo del

período de afloramiento. En consecuencia, desde finales de la década de

1980, se vienen realizando muestreos hidrográficos intensivos, tanto a

escala espacial, 1 estación cada 25 km2, como temporal, dos veces por

semana (Álvarez–Salgado et al. 1996a; 1996b; 2001; Rosón et al. 1999; Pérez

et al. 2000b; Gilcoto et al. 2001). Así, el balance de un compuesto C en la

capa superficial de la ría puede expresarse mediante la ecuación (Figura

16):

PNEATMCQCQ

)CC(MCQt

)CV(d

SSRR

SBZZZSS

++⋅−⋅+

+−⋅+⋅=∆

(5)

donde VS es el volumen de la capa superficial (en m3); ∆t es el tempo

transcurrido entre dos muestreos consecutivos (en s); CS y CB son las

concentraciones de C en las capas superficial y de fondo (en mmol m-3); CZ

es la concentración de C en el fondo de la capa superficial da ría

(en mmol m-3); CR es la concentración de C en el río (en mmol m-3); ATM

es el intercambio de C con la atmósfera, incluyendo la evaporación y la

precipitación en el caso de sustancias químicas disueltas y la difusión en el

caso de los gases (en mol s-1); y PNE es la producción neta de C en el

ecosistema (en mol s-1), es decir, la producción do fitoplancton menos la

respiración de todos los organismos que viven dentro de VS (Smith y

Hollibaugh 1997). Puede escribirse un balance similar para la capa de

Page 69: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

51

fondo, omitiendo los términos relativos a los intercambios de C con las

aguas continentales y con la atmósfera.

Los trabajos que presentan balances biogeoquímicos no estacionarios

aplicados en las rías gallegas hasta la solicitud del proyecto REMODA

(año 2000) se ciñen a la Ría de Arousa (Álvarez–Salgado et al. 1996a;

1996b; Rosón et al. 1999; Pérez et al. 2000b).

Otros estudios anteriores al proyecto REMODA es el trabajo de

Moncoiffé et al. (2000) que realizaron el primer estudio de acoplamiento a

corta escala espacio-temporal de la producción y respiración del

microplancton en la Ría de Vigo.

En estos trabajos se muestra que la PNE de materia orgánica durante el

período de afloramiento de 1989 en la Ría de Arousa fue de 150 mg m-2 d-1

(Figura 24) con una composición bioquímica media de C147H240O39N21P, es

decir, con un 53% de proteínas, 10% de compuestos de fósforo, 32% de

lípidos y 5% de carbohidratos que difiere sustancialmente de Redfield

(Tabla 2).

Page 70: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

52

sedimentos

DON65

PON

50

20

150100NEPR

M

15

d

SED

sedimentos

DONDON65

PONPON

50

20

150100NEPR

MM

15

d

SED

Figura 24. Balances medios de nitrógeno (en mg m-2 d-1) en la Ría de Arousa del 08/06 al 15/10/1989.) R, respiración; NEP, producción neta del ecosistema; DON, exportación de nitrógeno orgánico disuelto hacia la plataforma; PON, exportación de nitrógeno orgánico particulado hacia la plataforma; SED, exportación de nitrógeno orgánico particulado hacia los sedimentos de la ría; M, extracción de N orgánico particulado para consumo humano. Adaptado de Álvarez–Salgado y Fraga (en prensa).

La composición de la materia orgánica que se produce en las rías es

realmente muy variable, tanto en el espacio (de ría a ría; entre diferentes

partes de la misma ría; entre profundidades distintas en la misma

posición) como en el tiempo (períodos de afloramiento y hundimiento;

distintas fases del ciclo tensión/relajación del viento costero; ciclos

día/noche) en función de la disponibilidad de sales nutrientes y de las

características de la comunidad de organismos planctónicos dominante en

cada momento (Figueiras y Niell 1986; Fraga et al. 1992; Ríos 1992; Ríos et

al. 1998; Álvarez–Salgado et al. 1998).

A la escala temporal del período de afloramiento, sobre 1/3 de la PNE

(50 mg N m-3 d-1) se exporta hacia la plataforma en forma de nitrógeno

orgánico particulado (Figura 24). Extrapolando la razón carbono orgánico

disuelto/carbono orgánico particulado de 1.3 que obtuvieron,

posteriormente a la solicitud del proyecto REMODA, Gago et al. (2003) en

la Ría de Vigo durante el período de afloramiento de 1997, resultaría que

la Ría de Arousa exporta hacia la plataforma 65 mg N m-2 d-1 de nitrógeno

orgánico disuelto. Por otro lado, el cultivo intensivo de mejillón en la Ría

Page 71: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

53

de Arousa extrae unos 10–15 mg N m-2 d-1 (Álvarez–Salgado et al. 1996a).

Por consiguiente, entre 20 y 25 mg N m-2 d-1 se depositarán en los

sedimentos de la ría (13–17% da PNE).

La Figura 25 muestra la evolución temporal a corta escala de la PNE de

la Ría de Arousa durante el período de mayo a Octubre 1989. Se observa

una sucesión de pulsos de intensa PNE separados por períodos en los que

esta disminuye bruscamente hasta hacerse incluso negativa, es decir,

mostrando cortos episodios de heterotrofía neta (Figura 25d). Los pulsos

de producción/consumo de nitrógeno orgánico acompañan claramente a

los de tensión/relajación de los vientos favorables al afloramiento (Figura

25a). Además del viento en la plataforma, la estabilidad de la columna de

agua (Figura 25c) también contribuye a modular la PNE, al reducir a

difusión vertical de sales nutrientes desde la capa de fondo a la de

superficie durante los períodos de relajación y calma e incrementar la

eficiencia de la ría como trampa de sales nutrientes durante los episodios

de afloramiento. En este sentido, si la estratificación de la capa de

superficie no se rompe durante un episodio de afloramiento, la eficiencia

es mayor porque se fertiliza fitoplancton adaptado a las condiciones de

luz de la capa superficial. Por el contrario, si la estratificación se rompe, el

fitoplancton que aflora desde la capa de fondo sustituirá al que se

encontraba previamente en superficie, de modo que no está adaptado al

aumento brusco de intensidad de luz que experimenta al ascender,

necesitando un tiempo para adecuar el sistema fotosintético a esas novas

condiciones (Zimmerman et al. 1987).

Así, Pérez et al. (2000b), en un intento de predecir la PNE a partir do

viento de plataforma (EX), el intercambio de calor con la atmósfera (H) y la

Page 72: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

54

estabilidad de la columna de agua (BV), llegaron a la siguiente expresión

empírica:

0N50)·285( -)H (H 0,09)·0.55( E · 20)50( -25)·E100( PNE 1/2-0x01-x ><±+±+±±=

r = 0,84, n = 46, p < 0.0 (6)

donde Ex0 y Ex–1 son el transporte de Ekman medio entre los dos

muestreos consecutivos en los que se calcula la PNE y entre los dos

muestreos consecutivos de la semana anterior (en m3 s-1 km-1 de costa); H0

y H–1/2 son el intercambio de calor con la atmósfera entre los dos

muestreos consecutivos en los que se calcula la PNE y los dos muestreos

consecutivos anteriores, es decir, media semana antes (en cal m-2 d-1); y

<N>0 es la estabilidad de la columna de agua entre los dos muestreos

consecutivos en las que se calcula la PNE (en min–1). Por lo tanto, de

acuerdo con esta ecuación, el 70% de la variabilidad de la PNE a escala de

3–4 días puede explicarse por medio de una simple combinación lineal de

tres variables físicas. La ecuación nos indica que la intensidad del

afloramiento en la semana anterior (Ex–1) es la variable que más hace

aumentar la PNE. En contraste, el afloramiento durante el período de

muestreo (Ex0) reduce la PNE porque activa la circulación residual

diminuyendo el tiempo de residencia del fitoplancton que crece en la capa

superficial de la ría a expensas de las sales nutrientes que la fertilizaron la

semana anterior.

Page 73: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

55

mes de 1989

NEP

(mg

N m

-2 d-1

)

-200

0

200

400

600 <N>

(min

-1)

0.0

0.5

1.0

1.5

H (c

al c

m-2

d-1

)

-2000

200400600800

E X (1

03 m3 s-1

km

-1)

-4

-2

0

2

VZ

(m d

-1)

-60

-30

0

30

PNE predichaPNE calculada

a

d

c

b

OctubreSeptiembreAgostoJulioJunio

EXVZ

Figura 25. Evolución temporal del transporte de Ekman, Ex (a) el intercambio de calor con la atmósfera, H (b) la estabilidad de la columna de agua o frecuencia de Brunt-Väsisälä, <N> (c) y la producción neta del ecosistema, PNE (e) calculada por balances de materia (línea roja) y con la ecuación de Pérez et al. 2000b (línea azul) entre dos muestreos consecutivos en la Ría de Arousa do 08/06 al 30/10/1989. La estabilidad media de la columna de agua se calcula como

UL /ln()z/g2(N ρρ××>=< , donde g es la aceleración de la gravedad, z es la profundidad de la

columna de agua y ρU y ρL son las densidades medias de las capas de superficie y fondo de la ría entre dos muestreos consecutivos (Millard et al. 1990). Adaptado de Pérez et al. (2000b).

Page 74: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

56

Respecto de la síntesis de estructuras silíceas y calcáreas, la relación

N/Si de la PNE es de 3,2 mol N/mol Si (Pérez et al. 2000b). Las diatomeas

creciendo bajo condiciones en las que ni el N ni el Si son limitantes tienen

una relación N/Si media de 1,0 (Brezinkski 1985). Estas son las

condiciones que se dan en las rías durante el período de afloramiento; en

el caso concreto de la Ría de Arousa no se utilizó el 42% do nitrógeno

inorgánico ni el 51% del silicato que circulaba por la ría en el período de

afloramiento de 1989. Según esto, puesto que las diatomeas son la única

población de fitoplancton de la ría con capacidad de crear estructuras de

sílice biogénica (Pérez et al. 2000b; Tilstone et al. 2000), resulta que este

grupo taxonómico contribuiría a un 31% de la PNE. Por otra parte, los

mejillones de la Ría de Arousa producen 0,1 g C m-2 d-1 en forma de

CaCO3 (Rosón et al. 1999). Este valor no se diferencia significativamente

de la tasa de calcificación obtenida por Rosón et al. (1999) en base a un

balance da alcalinidad de la Ría de Arousa durante el período de

afloramiento de 1989.

En base a la ecuación estequiométrica de la PNE en la Ría de Arousa, la

PNE en unidades de carbono resulta ser 0,85 g C m-2 d-1 y la producción

media del fitoplancton (Pg) estimada por el método del 14C en el mismo

período fue de 1,50 g C m-2 d-1 o 250 mg N m-2 d-1 (Álvarez–Salgado et al.

1996a). La diferencia entre PP y PNE nos da una respiración media del

ecosistema (R) de 0,65 g C m-2 d-1 o 100 mg N m-2 d-1, es decir, un 40% de

la PP se respira en la columna de agua y un 60% queda disponible para 1)

transferirse a niveles tróficos superiores, esencialmente al mejillón de

batea (6% de Pg); 2) exportarse a la plataforma adyacente en forma de

materia orgánica disuelta (26% de Pg) y particulada (20% de Pg) y 3)

preservarse en los sedimentos de la ría (8% de Pg).

Page 75: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

57

Por último, en una situación de calma prolongada (más de una semana)

la Pg disminuye drásticamente a consecuencia del agotamiento de las

sales nutrientes y la R aumenta a expensas de la materia orgánica

producida durante el anterior período productivo, donde R es mayor que

PP, dando lugar a una situación transitoria de heterotrofía neta en la que

predomina la “red trófica microbiana”.

Al contrario de lo que sucede en el período de afloramiento, la

información existente sobre los procesos biogeoquímicos que operan en

las rías a escala temporal de 3–4 días durante el período de hundimiento

es escasa, concentrándose fundamentalmente en Septiembre–Octubre

porque es la época en la que suelen ocurrir las “mareas rojas”. Las Figuras

18 y 26 muestran la circulación, los flujos y el balance biogeoquímico del

nitrógeno durante un episodio de hundimiento en la Ría de Arousa en el

mes de Octubre de 1989. Las situaciones de hundimiento otoñal se

caracterizan por heterotrofía neta (R > PP), fundamentalmente a expensas

de material importado desde la plataforma.

Parte de las sales de nitrógeno que fertilizan las rías vienen de la

mineralización del nitrógeno orgánico particulado que se exporta desde

las propias rías, de forma que se consigue una gran eficacia en la

utilización de las sales nutrientes en el sistema formado por las rías y la

plataforma (Fraga 1981; Tenore et al. 1982; S Fraga y Bakun 1993; Álvarez–

Salgado et al. 1993; Prego 1994). Álvarez–Salgado et al. (1997) observaron

que el enriquecimiento en sales nutrientes de la ACNAE aflorada sobre la

plataforma aumenta desde el cantil continental hasta la boca de las rías,

reproduciendo el patrón que se observa en la distribución del contenido

de materia orgánica (López–Jamar et al. 1992) y sílice biogénica (Prego y

Bao 1997) en los sedimentos.

Page 76: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

58

En la Figura 26 se hace una propuesta de balance del nitrógeno en la

Ría de Arousa durante el episodio de hundimiento de Octubre de 1989, en

el que se sugiere que de los 146 mg N m-2 d-1 de nitrógeno orgánico que se

mineralizan en la ría sólo 78 mg N m-2 d-1 quedan en forma de nitrógeno

inorgánico disuelto en el agua. Los 68 mg N m-2 d-1 restantes (un 47%) se

pierden a la atmósfera en forma de N2 utilizando un modelo

estequiométrico que incluye procesos de mineralización anaerobia para

justificar las razones molares O2/C/N/P observadas (Álvarez-Salgado y

Fraga en prensa).

PON-53

sedimento

68N2

2

NT78

d

DON-23

PON-53

sedimento

68N2 68N2

2

NT78

d

DON-23

Figura 26. Balances medios de nitrógeno (en mg m-2 d-1) en la Ría de Arousa del 15/10 al 30/10/1989. NT, nitrógeno inorgánico; DON, nitrógeno orgánico disuelto; PON, nitrógeno orgánico particulado; N2, nitrógeno molecular. Adaptado de Álvarez–Salgado y Fraga (en prensa).

La información existente sobre medida de procesos biogeoquímicos es

aún más escasa en invierno. Únicamente se ha estudiado el balance

biogeoquímico de las rías a escala temporal de 3–4 días posteriormente a

la solicitud del proyecto REMODA en: Álvarez–Salgado et al. (2001) y en

el Capítulo 4 de esta memoria (ambas en la Ría de Vigo). Álvarez–Salgado

et al. (2001) evaluaron los flujos y el balance biogeoquímico del carbono

orgánico disuelto (DOC) obteniéndose que el intercambio de DOC entre la

ría y la plataforma adyacente se deben más a las variaciones en el término

Page 77: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

59

de acumulación (∆DOC /∆t) que a la producción o consumo in situ de

DOC.

Si bien no se puede establecer un patrón de comportamiento de las rías

en invierno en base a las pocas observaciones existentes, en general puede

decirse que la Pg calculada por métodos in vitro suele estar por debajo de

los 0,5 g C m-2 d-1 y, ocasionalmente, se han medido valores menores de

0,05 g C m-2 d-1 (Fraga 1976; Bode y Varela 1998). Se trata pues de un

período de baja actividad del fitoplancton.

Posteriormente a la solicitud del proyecto REMODA surgieron otros

trabajos de modelos biogeoquímicos cuyo objeto de estudio era la Ría de

Vigo. Entre ellos están el de Míguez et al. (2001a) que relaciona las

condiciones hidrodinámicas de la Ría de Vigo, resultado de la aplicación

de un modelo de cajas no estacionario, con la sucesión fitoplanctónica

asociada a lo cambios en las condiciones oceanográficas. El modelo de

cajas desarrollado por Gilcoto et al. (2001) con la inclusión de los métodos

inversos fue aplicado también a la Ría de Vigo, y permitió a su vez

calcular la NEP. Este modelo de cajas fue aplicado posteriormente por

Gago et al. (2003) y Álvarez-Salgado y Gilcoto (2004) en la Ría de Vigo,

para estudiar el destino del carbono y los procesos de nitrificación,

respectivamente.

Además de las aportaciones de los modelos biogeoquímicos, la

contribución de trabajos a las rías gallegas que incorporan medidas de

producción y respiración ha sido muy prolífica durante los últimos cinco

años, atendiendo a diversos aspectos, ya sea la composición química, la

producción o el destino de la materia orgánica producida (Teira et al.

2003; Bode et al. 2004; Marañón et al. 2004; Varela et al. 2004; Nieto Cid et

al. 2004; 2005; 2006, Tilstone et al. 2003; Cermeño et al. 2005; Lorenzo et al.

Page 78: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

60

2005). Un análisis global de todas estas estudios concluye que la

producción primaria es muy variable durante el período de afloramiento,

con valores que van desde menos de 0,5 hasta más de 10 g C m-2 d-1, con

un valor medio de 2,5±2.8 g C m-2 d-1 (Arístegui et al. 2006; Álvarez–

Salgado et al. 2006 y las referencias citadas en estos trabajos).

Sin embargo, hay una carencia de trabajos tanto en las rías como en la

plataforma adyacente que incluyan medidas de tasas de sedimentación y

de herbivoría. Las tasas de sedimentación de material biogénico en las rías

son poco conocidas. De hecho, sólo se dispone de un trabajo en el que esas

tasas se obtuvieron de forma directa usando trampas de sedimentación en

la Ría de Pontevedra (Varela et al. 2004). Estos autores obtuvieron que la

exportación media diaria de carbono orgánico biogénico desde la capa

superficial a la capa de fondo fue de un 75% de la producción primaria

(PP), con valores mínimos durante los episodios de afloramiento, cuando

domina la exportación horizontal, y máximos durante los episodios de

hundimiento.

En cuanto a trabajos que incluyan tasas de herbivoría en las rías la

carencia es total. La única aportación es la de Fileman et al. (2001) en

plataforma.

Page 79: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

61

6. Hipótesis general

La circulación tridimensional en dos capas en el segmento central de la

Ría de Vigo está controlada por forzamientos externos, fundamentalmente

el viento en plataforma que actúa a escala instantánea, la circulación

gravitacional, y la topografía/geometría de la ría.

Este esquema de circulación tridimensional determina los flujos

biogeoquímicos por influir tanto en la variación de las propiedades cuyos

flujos se estiman como en la del propio flujo, condicionando el origen y el

destino del material biogénico producido ya sea por relaciones tróficas (ej.

respiración y herbivoría) o por trasporte (ej. sedimentación y exportación

horizontal).

7. Objetivos

1) Caracterizar la estructura vertical de la circulación residual en el

segmento central de la Ría de Vigo en cuatro escenarios oceanográficos,

mediante:

- El análisis cuantitativo de la importancia relativa de los diferentes

forzamientos que afectan al sistema (viento costero, viento local, aportes

continentales e intercambio de calor con la atmósfera).

- El análisis del tiempo de respuesta de la columna de agua a los

forzamientos externos.

- El análisis del tipo de circulación existente bajo condiciones

atmosféricas e hidrográficas diferentes.

- El análisis de la profundidad del “nivel de no movimiento” entre

capas que fluyen en direcciones opuestas bajo condiciones atmosféricas e

hidrográficas diferentes.

Page 80: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

62

2) Caracterizar la variabilidad lateral de la circulación y las variables

termohalinas y químicas en el segmento central de la Ría de Vigo, bajo

escenarios meteorológicos, hidrográficos y dinámicos diferentes,

mediante:

- El análisis de las diferencias laterales en las componentes longitudinal

y transversal de la corriente mediante la combinación de medidas de

corrientes obtenidas a partir del fondeo de correntímetros y a bordo del

B/O Mytilus y los resultados del modelo numérico HAMSOM.

- El análisis de las diferencias laterales en las variables termohalinas y

químicas.

3) El estudio del origen y destino de un “bloom” de la diatomea

Skeletonema costatum que ocurrió en la ría durante un periodo de 10 días en

Febrero de 2002, mediante la combinación de datos meteorológicos,

hidrodinámicos e hidrográficos (termohalinos, químicos y biológicos) y

las medidas obtenidas de experimentos de incubación (medidas in vitro

de producción, respiración y herbivoría) y un balance de materia.

4) La estimación de los flujos de compuestos inorgánicos y orgánicos a

partir de las medidas de corriente para evaluar el estado trófico del

sistema a través de un balance geoquímico (in situ) en dos escenarios

hidrográficos: afloramiento/relajación estival, hundimiento/relajación

otoñal.

5) La comparación del balance de la materia orgánica in situ con un

balance de materia orgánica in vitro, a partir de la estimación de tasas de

producción primaria/respiración de materia orgánica por el

microplancton marino, en los dos escenarios hidrográficos anteriores. Para

ello se evaluará también la intervención del fitoplancton,

microheterótrofos (bacterias y protozoos) en la utilización de la materia

Page 81: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Introduction

63

biogénica que se produce dentro del sistema, y la importancia de la

sedimentación del Material Orgánico Particulado (MOP).

8. Estructura

Esta memoria consta de 8 capítulos:

- El capítulo 1 es una presentación del sistema de estudio, en la que se

describen las características geomorfológicas, meteorológicas y dinámicas

de las Rías Baixas, haciendo hincapié en el estado de conocimiento de los

procesos hidrodinámicos, biogeoquímicos y del acoplamiento entre ambos

en el momento en el que se solicitó el proyecto REMODA.

- Los capítulos 2 y 3 describen a los procesos físicos que tienen lugar el

sistema, analizando la variabilidad temporal de corta escala y la

variabilidad espacial en cuatro escenarios oceanográficos.

- Los capítulos 4, 5 y 6 combinan datos meteorológicos, hidrodinámicos

e hidrográficos (termohalinos, químicos y biológicos) así como las

medidas obtenidas de experimentos de incubación (producción,

respiración, herbivoría y sedimentación), para realizar una caracterización

biogeoquímica de tres escenarios oceanográficos.

- El capítulo 7 resume las conclusiones más relevantes de la memoria.

- El capítulo 8 muestra la literatura citada.

Page 82: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 1

64

Page 83: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

 

                         

            Chapter 2: 

Short-timescale thermohaline variability and

residual circulation in the central segment of the

coastal upwelling system of the Ría de Vigo

Page 84: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 85: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

67

Chapter 2, the research work presented in this chapter is also a

contribution to the paper:

S. Piedracoba, X. A. Álvarez-Salgado, G. Rosón, J. L. Herrera. 2005.

Short-timescale thermohaline variability and residual circulation in the

central segment of the coastal upwelling system of the Ría de Vigo

(northwest Spain) during four contrasting periods. J. Geophys. Res. 110

(C03018), 1-15, doi:10.1029/2004JC002556.

Abstract: Sixteen hydrographic surveys were carried out in the middle

segment of the Ría de Vigo (northwest Spain) at 3- to 4-day intervals

during February, April, July, and September 2002 (four surveys per

period). Simultaneously, an acoustic Doppler current profiler (ADCP)

mooring recorded the velocity profile. Combination of direct current

measurements with the output of an inverse model based on the time

course of the distributions of salinity and temperature allowed an

objective analysis of the effect of the meteorological forces on the

hydrodynamics of the Ría. Remote shelf winds explained more than 65%

of the variability of the subtidal circulation, which responded immediately

to this forcing (lag time, <2 days). Shelf winds created a simple two-

layered circulation pattern, with a surface outgoing current under

northerly winds and a surface ingoing current under southerly winds. A

three-layered circulation developed during the transitions from northerly

to southerly winds and vice versa. At the same time, an empirical

orthogonal function (EOF) analysis demonstrated the lack of contribution

of local winds to the subtidal dynamics of the ría. Continental runoff and

heat exchange with the atmosphere explained less than 5% and 25% of the

variability observed in the subtidal circulation of the Ría de Vigo.

Page 86: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

68

Page 87: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

69

Resumen: Se realizaron 16 muestreos hidrográficos en el segmento

central de la Ría de Vigo (NO de España) con una frecuencia de 3-4 días

durante los meses de Febrero, Abril, Julio y Septiembre de 2002

(4 muestreos por periodo). Simultáneamente se registró el perfil de

velocidades mediante el fondeo de un perfilador de corrientes de efecto

Doppler (ADCP). La combinación de medidas directas de corriente,

conjuntamente con los resultados de un modelo inverso desarrollado a

partir de la evolución temporal de las distribuciones de salinidad y

temperatura permitieron evaluar el efecto de los forzamientos

meteorológicos en la hidrodinámica de la ría. Se concluye que el viento de

plataforma explica más del 65% de la variabilidad de la circulación

residual, que responde de forma inmediata a este forzamiento con un

desfase <2 días. El viento de plataforma da lugar a un modelo de

circulación en dos capas con una corriente superficial de salida bajo la

acción de vientos del Norte y por el contrario, una corriente superficial de

entrada bajo la acción de vientos del Sur. Transiciones del viento Norte-

Sur y Sur-Norte permiten el desarrollo de periodos cortos de circulación

en tres capas.

Análogamente, la aplicación de funciones empíricas ortogonales (EOF)

demuestra la falta de contribución del viento local a la dinámica residual

de la ría. Por otra parte, los aportes de agua continental y el intercambio

de calor con la atmósfera explican menos del 5% y 25% de la circulación

residual de la Ría de Vigo, respectivamente.

Page 88: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

70

Page 89: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

71

1. Introduction

Coastal embayments and estuaries respond to a variety of forcing

mechanisms over a wide range of timescales (Wong and Moses-Hall 1998).

Subtidal motion ultimately determines the long-term transport of

suspended and dissolved materials, even though the semidiurnal or

diurnal tidal motions are often the most energetic mechanisms operating

in estuaries. Regarding subtidal motion, the density-induced gravitational

circulation was the first to attract extensive research (Pritchard 1952, 1956).

During the 1960s and 1970s, the two-layered gravitational circulation was

established as the basic residual flow pattern associated with partially

mixed estuaries.

The importance of the atmospheric forcing for the subtidal variability

was not fully recognized until the late 1970s, when Carter et al. (1979)

reviewed the dynamics of motion in estuaries made by U.S. researchers

during the period 1975–1978. This review highlighted that wind-induced

fluctuations dominated the subtidal circulation of the Providence River

(Weisberg and Sturges 1976), the west passage of Narragansett Bay

(Weisberg 1976) or the Chesapeake Bay (Wang and Elliott 1978).

Especially relevant was the paper by Elliott (1978) showing that 25% of the

total variability of the residual currents in the Potomac Estuary was

associated with remote winds from the adjacent shelf. The scientific

interest on the wind-induced subtidal variability increased during the

1980s, with process-orientated studies conducted in estuaries such as San

Francisco Bay (Walters 1982), Delaware Bay (Wong and Garvine 1984),

and Mobile Bay (Schroeder and Wiseman 1986) or some coastal lagoons

with restricted communication with the ocean (e.g. Wong and Wilson

1984). Results from the above-mentioned studies revealed that the remote

Page 90: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

72

atmospheric forcing could produce subtidal variability in estuaries and

coastal inlets by the impingement of coastal sea level fluctuations at the

mouth of the estuary (Wong and Moses-Hall 1998).

The Rías Baixas (northwest Spain) are four large (>2.5 km3) and V-

shaped coastal inlets, freely connected with the adjacent shelf. The ría

produce annually more than 250,000 tons of mussels, which represent 25%

of the world production (Figueiras et al. 2002). The nutrients necessary to

support this culture are supplied by upwelled Eastern North Atlantic

Central Water (ENACW), which eventually enters the rías (Fraga 1981;

Rosón et al. 1997; Álvarez-Salgado et al. 2000). The western coast of the

Iberian Peninsula is located at the boundary between the temperate and

subpolar regimes of the coastal upwelling system of the Eastern North

Atlantic, where upwelling-favorable northerly winds predominate from

March–April to September–October in response to the seasonal migration

of the Azores High.

On the contrary, downwelling-favorable southerly winds prevail the

rest of the year (Wooster et al. 1976; Bakun and Nelson 1991). Warm and

salty surface waters of subtropical origin pile on the shelf during

downwelling events (Haynes and Barton 1990, 1991). However, this

pattern explains <20% of the variability observed, whereas >70%

concentrates at frequencies <30 days (Nogueira et al. 1997).

The recognized importance of coastal upwelling for the productivity of

the rías focused the research during the 1990s on the effect of remote shelf

winds on the hydrography of these coastal inlets at the timescale of an

upwelling episode, 1–2 weeks (Álvarez-Salgado et al. 1993). Inverse

modeling of thermohaline data collected with a one-half-week periodicity

has been a useful tool to study the coupling between the wind blowing in

Page 91: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

73

the adjacent shelf and the subtidal circulation of the Ría de Arousa (Rosón

et al. 1997), Ría de Vigo (Álvarez-Salgado et al. 2000; Gilcoto et al. 2001)

and Ría de Pontevedra (Pardo et al. 2001). These studies concluded that

shelf winds were mainly responsible for the changes observed in the

hydrography of the rías and that these changes were compatible with the

assumption of a two layered residual circulation pattern driven by shelf

winds.

A new tool has been recently introduced for the study of the remote

wind-induced circulation of the rías: Torres-López et al. (2001) and Souto

et al. (2001, 2003) applied the GHER (GeoHydrodynamics and

Environment Research) and HAMSOM (HAMburg Shelf Ocean Model)

numerical models, respectively. They compared simulated with

experimental data to conclude that the typical estuarine two layered

circulation pattern of the inner and middle Ría de Vigo overlaps with the

alongshore circulation of the ría, which results from the interaction of

shelf winds and the intricate topography of the coast. Thus a three-

dimensional (3-D) residual circulation pattern has been proposed for the

outer part of the ría.

Despite the effort to understand the dynamics of the Rías Baixas, and

particularly the Ría de Vigo, direct current measurements are scarce, with

a restricted spatial and temporal coverage. Álvarez-Salgado et al. (1998a)

described an abnormal three-layered circulation pattern during a 1-day

sampling in the middle Ría de Vigo in September 1991. Gilcoto et al. (2001)

showed the coupling between the measured surface circulation of the

middle ría and shelf winds during 20 days in September 1990. Finally,

Míguez et al. (2001b) and Souto et al. (2003) focused on the relationship

between the residual currents at the northern and southern mouths of the

Page 92: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

74

outer Ría de Vigo and the winds blowing over the shelf. They were able to

compute lag times ranging from 0 to 1 days at both mouths. However,

until now it was unknown what is the relative importance of the different

forcing to the residual circulation, what is the lag time of the water

column response to the dominant agent forcing in a central segment of the

ría, and in what conditions can be developed a scheme of circulation in

three layers.

With the aim of amending the lack of repeated current meter records

under contrasting atmospheric and hydrographic conditions, an intensive

program was conducted in the coastal upwelling system of the Ría de

Vigo during 2002. Northerly and southerly wind conditions of variable

intensity, as well as water column stratification (thermal and haline) and

mixing conditions, were captured.

In addition, direct current meter data were combined with a 2-D

inverse model to provide complementary views of the residual circulation

of the ría. For the first time, the vertical structure of the residual

circulation of the middle ría will be resolved for a wide variety of

atmospheric and hydrographic conditions, focusing on (1) the quantitative

assessment of the relative importance of the different external forces

acting on the system (remote and local winds, continental runoff, and heat

exchange with the atmosphere) and (2) the lag time of the water column

response in a well-protected site (15 km inshore of the shelf), compared

with the outer ría (directly exposed to shelf winds). The contrasting

conditions for (3) the development of the two-layered or the three-layered

circulation patterns observed in the middle ría and (4) the depth of the

level of no motion (LNM) separating layers flowing in opposite directions,

will also be described. Finally, (5) the asymmetry in the response of the

Page 93: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

75

central segment of the Ría de Vigo to shelf winds will be examined on the

basis of the differences observed in the salinity and temperature of a

transverse hydrographic section. The results will be discussed in the

wider context of the response of estuaries and coastal inlets to remote and

local forcing agents.

2. Materials and Methods

Sixteen surveys were carried out aboard R/V Mytilus during four

contrasting periods in 2002 (four surveys per period). Every survey, a

CTD SBE-25 was dipped at five selected stations along a transect

transversal to the main axis of the embayment, from Cabo de Mar to

Punta Borneira (Figure 1). An acoustic Doppler current profiler (ADCP)

was moored at station R00, in 40 m water, to obtain a continuous record of

the velocity profile during the four periods. In addition, a series of

conductivity temperature-depth (CTD) profiles along the main axis of the

ría, from San Simon Bay to station R00, collected during 61 surveys to the

Ría de Vigo between 1990 and 1997 were also used to complement the

hydrographic data collected during 2002.

2.1. Measured Variables

Current velocity profiles were measured with an Aandera DCM12 at

the central station R00 (42º14.071’ N and 8º47.208’ W). The site was chosen

because it has been proven suitable for evaluating the main processes that

take place in the system due to changes in the external forcing factors

(Nogueira et al. 1997). The mooring was deployed during 24 days in

February, 10 days in April, 19 days in July, and 9 days in September 2002.

The DCM12 Doppler Current Meter was deployed on the seabed, and the

following parameters were recorded at 30-min intervals: (1) the current at

the sea surface and five depths and (2) the water level and significant

Page 94: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

76

wave height (H1/3). Cell size was 11 m, centered at 6.5, 13, 19.5, 26, and

32.5 m, corresponding to L1, L2, L3, L4, and L5, respectively.

Figure 1. (a) Map of the Ría de Vigo showing the land based meteorological stations of Bouzas and Peinador and the Eiras reservoir. (b) Section along the main channel of the Ría de Vigo showing the study volume used in the inverse box model application, with surface and bottom layer. The outer limit is the section between Cabo de Mar and Punta Borneira; QR , river discharge; E, evaporation rate; P, precipitation; H, atmospheric heat flux; QS and QB are the surface and bottom horizontal advective fluxes measured at the outer boundary. (c) Bathymetry of the Ría de Vigo and the adjacent shelf showing the meteorological station off Cape Silleiro and a detailed cross section between Cabo de Mar and Punta Borneira including stn R00, where the Aanderaa DCM12 Doppler current meter was moored.

Other data used in this study were (3) local winds, from the ship-

mounted meteorological station and the meteorological observatory of

Bouzas, National Institute of Meteorology (Figure 1); (4) remote winds,

from the Seawatch buoy of Puertos del Estado off Cape Silleiro, a

representative site for winds blowing off the Ría de Vigo (Herrera et al.

2005) ; (5) daily precipitation rates, from the meteorological observatory of

Page 95: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

77

Peinador airport, National Institute of Meteorology (Figure 1); and (6)

humidity, air temperature, and atmospheric pressure, from the

meteorological observatory of Bouzas. A A242A25 filter with a cutoff period

of 30 hours was passed to the time series of winds and currents, to remove

the variability at tidal or higher frequencies (Godin 1972). 2.2. Variables calculated from collected data

Daily values of the offshore Ekman transport (–QX, m3 s–1 km–1) were

calculated from Wooster et al. (1976):

CSW

yDairX f

WWCQ

ρ

ρ−=− (1)

Where ρair is the density of air, 1.22 kg m–3 at 15 ºC. CD is an empirical

drag coefficient (dimensionless), 1.3 10–3 according to Hidy (1972), fC is the

Coriolis parameter at 42º latitude. ρSW is the density of seawater, ~1025 kg

m–3. W and Wy are the wind speed and the north component of wind

speed recorded at the Seawatch buoy off Cape Silleiro.

Continental runoff to the inner Ría de Vigo is a combination of

regulated and natural flows. The Eiras reservoir controlled 42±16 % of the

total flow of the River Oitavén–Verdugo during the study year 2002. Daily

flows were provided by the company in charge of the management of

urban waters. The natural component of the flow per unit area (QR/A, in l

m–2 d–1) was calculated according to the empirical equation of Ríos et al.

(1992b) from the daily precipitation in the drainage basin, P(n):

n30

1n130

R knPkk

k1A

Q )(∑=

+−−

= (2)

This equation accounted for the influence of precipitation during

the 30 days before the study date. The retention constant k, has a value of

0.75 for the 586 km2 drainage basin of the Ría de Vigo (Ríos et al. 1992b).

Page 96: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

78

River discharge data were not filtered, as they came in the form of daily

average values.

Evaporation rates (E, mm d–1) were calculated with the empirical

equation of Otto (1975), based on the local wind velocity (W, m s–1) and

the vapour pressure at the sea surface (es, in mbar) and 2 m above the sea

surface (ez):

))(..( zs eeW0770260E −+= (3)

100Hu)s000537.01(

))T1065.4(P101(ee 2A

3r

6Tz A

××−×

×××+××+×= −−

(4)

)s000537.01(

))T1065.4(P101(ee 2S

3r

6Ts S

×−×

×××+××+×= −−

(5)

Where s (in psu), TA and TS (in ºC) are the surface salinity and the air

and water temperatures, Pr is the atmospheric pressure and Hu is the

relative humidity. ATe and

STe (in mbar) are the distilled water vapour

pressure at AT and ST , which can be calculated for temperatures between

–40 and 40 ºC with (Gill 1982):

T00412.01T03477.07859.0elog T ×+

×+= (6)

Heat exchange with the atmosphere (H) was evaluated considering the

balance of the following terms: irradiation, conduction, back radiation,

reflection and heat lost by evaporation. Irradiation (I, cal cm–2 d–1) was

calculated with the equation developed by Rosón et al. (1997) for 42ºN:

( )N474.4417.50365

J355sin15.1119.3I 2 −×⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛ −π+= (7)

J is Julian day, from 1 (1 January) to 365 (31 December) and N is the

cloudiness in oktas. Conduction (C, in cal cm–2 d–1) was obtained with the

Page 97: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

79

empirical equation of Otto (1975) that depends on the temperature

gradient between the sea surface ( ST , in ºC) and the atmosphere ( AT ),

and on local wind speed (W, in m s–1):

)TT)(W114.038.0(88.24C AS −+= (8)

The back radiation term (B, cal cm–2 d–1) was estimated with the

equation of Laevastu (1963):

)N1.01)(H96.0T87.1297(B S −−−= (9)

Heat lost by reflection (R) was assumed to represent 6% of irradiation

at our latitudes (Otto 1975). Finally, evaporation also implies a loss of

energy that was calculated multiplying the rate of evaporation E (in mm

d–1) by 58.7 cal cm–2 mm–1.

Finally, the total heat balance was:

RBCEIH −−−−= (10)

2.3. Estimation of water flows with an inverse method

The 2–D, non–steady–state, salinity–temperature weighted box model

successfully applied by Álvarez–Salgado et al. (2000) to the Ría de Vigo

was used in this work. This box model is able to estimate the average

water fluxes (horizontal and vertical advection), under the assumptions of

volume, heat, and salt conservation. Horizontal advection fluxes were

calculated for the volume of the ría delimited by the transect from Cabo

de Mar to Punta Borneira between two consecutive surveys (Figure 1c).

A system of 3 linear equations ―conservation of water (11), salt (12)

and heat (13)― can be written for the study volume:

RQQ BS += (11)

SSBB SQSQtSV ⋅−⋅=∆∆ (12)

Page 98: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

80

HTRTQTQtTV RSSBB +⋅+⋅−⋅=∆∆ (13)

Where SQ and BQ (m3 s–1) are the average residual surface and bottom

horizontal fluxes across the outer boundary between two consecutive

surveys. The boundary is divided in two layers (surface and bottom)

flowing in opposite directions. The limit between the surface and bottom

layer (level of no–horizontal motion, LNM) was set at the pycnocline.

R (m3 s–1) is the freshwater balance of continental runoff ( RQ ),

precipitation ( P ) and evaporation ( E ). BT , ST and RT (ºC) are the

average temperature of the surface and bottom flows across the open

boundary of the box and the river flow, respectively. BS and SS are the

average salinity of the surface and bottom flows across the open boundary.

V is the volume of the box. H (m3 ºC s–1) is the average heat exchange

with the atmosphere. tS ∆∆ / and tT ∆∆ / are the net rate of change in

salinity and heat content (temperature) of the study volume between two

consecutive surveys.

Two sets of horizontal bottom convective flows are obtained from the

equations of water (11) and salt conservation (12), SBQ )( , and the

equations of water (11) and heat conservation (13), TBQ )( :

SB

S

SB SStSVSR

)Q(−

∆∆

+= (14)

SB

RS

TB TT

HtTV)TT(R

)Q(−

−∆∆

+−= (15)

A salinity–temperature weighted horizontal bottom flow, BQ , can be

obtained:

Page 99: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

81

)f1()Q(f)Q(Q TBSBB −+= (16)

Where the dimensionless factor, f , that weights the contribution of

salinity and temperature to the density gradient at the open boundary, is

calculated as:

22SB

2SB

2SB

)//()TT()SS()SS(

fαβ−+−

−= (17)

The coefficient αβ / converted the temperature gradient into salinity

units, where β and α are the coefficients of haline contraction and

thermal expansion, respectively.

The robustness of the estimated fluxes is inversely proportional to the

vertical gradients of salinity and temperature. This is the reason why the

factor f , was introduced to weight the relative contribution of the

salt, SBQ )( , and temperature, TBQ )( , solutions to the optimum salt–heat

weighted solution, TSBQ ,)( . During the winter months the value of f is

close to 1.0. Therefore, the optimum solution coincided with the solution

obtained from the salt balance because the temperature gradient is quasi

homogeneous, this is SB TT − = 0. In contrast, during the summer months

the value of f is close to 0.0. Therefore, the optimum solution coincided

with the solution obtained from the heat balance because the salt gradient

is quasi homogeneous, this is SB SS − = 0 .

Since tS ∆∆ / and tT ∆∆ / in the study volume were not measured, a

set of historical data for the same volume collected during the period

1990–1997 were used to infer the salinity and temperature of the box from

the salinity and temperature of the open boundary. The following linear

regressions were obtained:

Page 100: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

82

00R00RRÍA R)001.0(006.0S)05.0(2.1)2(8S ±−±+±−=

950R 2 .= , n = 61, p<0.0001 (18)

00RRÍA T)003.0(028.1T ±=

930R 2 .= , n = 61, p<0.0001 (19)

where 00RS , 00RT and 00RR are the salinity, temperature and hydrological

balance at stn R00 from the set of historical data. Therefore, tS ∆∆ / and

tT ∆∆ / in the inner ría were obtained from eqs (18) and (19), using the

thermohaline data recorded during the 2002 surveys in stn R00. Eq (16)

can be split in a steady–state term and a non–steady–term:

NSSBSSBB )Q()Q()Q( += (20)

Furthermore, the steady–state term, SSBQ )( , can be divided in a

hydrological term, HSSBQ ))(( , and a radiative term, RSSBQ ))(( :

)f1()TT()TT(R

fSS

SR))Q((SB

RS

SB

SHSSB −

−−

+−

= (21)

)f1()TT(

H))Q((SB

RSSB −−

−= (22)

2.4. Empirical orthogonal functions (EOF) analysis of local winds

An EOF analysis is a multivariate statistical technique that decomposes

the total variability of a time series into a set of modes ordered by the

percentage of the total variance that they explain. These modes are the

orthogonal and uncorrelated eigenvectors of the complex covariance

matrix of the time series and the eigenvalue of each eigenvector is the

percentage of the total variance explained by each mode (Kundu and

Allen 1976; Kelly et al. 1988).

Page 101: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

83

An EOF analysis of the February and July 2002 time series of Silleiro

and Bouzas winds and the surface (or bottom) flows at stn R00 was

preformed to study the relative influence of remote and local winds on the

dynamics of the ría. The differences in roughness length between the

remote (ocean based) and local (land based) meteorological stations were

accounted (Agsterberg et al. 1989) by normalising the friction between the

two locations and using the method of Wieringa (1986) to interpolate the

wind field originated from data taken at different locations.

2.5. Cross correlation between shelf winds and residual currents

The cross correlation coefficient is a statistical parameter that allows to

calculate the correlation between two time series W(t) and V(t+τ) that are

out phase with a lag time τ . The curve shape )(τR versus τ represents

the variation of the correlation between the two time series depending on

the lag time. This analysis was applied to the residual currents in the

middle ría along the preferent direction (the main axis of the ría) and the

North component of the wind recorded off Cape Silleiro during the 4

study periods.

3. Results

3.1 Local and remote winds

The results of the EOF analyses of the time series of subtidal surface

and bottom flows calculated from local winds, remote winds and ADCP

current measurements yielded similar results for February and July (Table

1). The analysis including the bottom flows calculated with the ADCP

current meter produced 2 main modes that explained 87% and 12% of the

total variability. The flows estimated from local winds do not contribute

significantly to both eigenvectors. On the contrary, the flows estimated

Page 102: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

84

from shelf winds and the flows estimated with the ADCP current meter

presented similar absolute values.

When the surface flows are included in the analysis, there was only one

dominant mode that explained 95% of the total variability. The flows

estimated from local winds had again a small contribution to the

eigenvector. However, the flows estimated from shelf winds yielded a

coefficient larger than the flows estimated from surface ADCP

measurements. This discrepancy probably arises from the lack of ADCP

measurements in the top few meters, where the current velocities are

larger. The small contribution of surface and bottom flows estimated from

local winds to the main EOF modes analyses is interpreted as a non

significant influence of local winds to the residual dynamics of the ría.

Since the EOF analyses revealed that only shelf winds have an influence

on the subtidal circulation of the Ría de Vigo, the wind regime recorded at

the Seawatch buoy off Cape Silleiro will be described for the winter,

spring, summer and autumn periods studied in 2002.

Figure 2a shows the low–pass filtered fluctuations of shelf winds

during February 2002, when two northerly wind episodes, which lasted

about 1 wk, separated by a short Southerly wind event occurred. The

intensity of the northerly winds ranged form –7 to –10 m s–1 during the

first episode and from –6 to –9 m s–1 during the second episode. Southerly

winds in between ranged from 6 to 9 m s–1. During April 2002, northerly

winds ranging from –3 to –6 m s–1 prevailed, but they persist for no longer

than 3 days (Figure 3a). The wind speed reduced to near –1 m s–1 during

the relaxation periods.

Page 103: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

85

EO

F m

ode

1 EO

F m

ode

2 EO

F m

ode

3

Febr

ury

July

Fe

brua

ry

July

Fe

brua

ry

July

S B

S B

S B

S B

S B

S B

%

95

87

97

88

3.4

12

2.9

12

1.3

0.7

0.1

0.1

Bouz

as

0.1

–0.0

9 –0

.01

–0.0

1 –0

.48

–0.0

7 0

–0.0

1 0.

9 1

–1.0

1

Sille

iro

–0.9

5 –0

.67

–0.9

9 –0

.81

0.3

0.7

–0.1

4 0.

6 0.

1 –0

.01

0 –0

.01

Mea

sure

d –0

.30

0.7

–0.1

4 0.

6 –0

.82

0.7

1 0.

8 –0

.49

0.1

0 0

During July 2002, northerly winds blew continuously but with a wide

range of intensities (Figure 4a). Three periods can be defined: a first

period, when northerly winds varied between –1 and –10 m s–1; a second

period of calm, with shelf winds intensities between 0 and –3 m s–1; and a

Tabl

e 1.

Eig

enve

ctor

s an

d pe

rcen

tage

of t

he to

tal v

aria

bilit

y of

the

time

seri

es o

f loc

al (B

ouza

s) a

nd s

helf

(Cab

o Si

lleir

o)w

inds

and

the

mea

sure

d su

btid

al fl

ow in

the

surf

ace

(S) a

nd b

otto

m (B

) lay

er o

f the

mid

dle

segm

ent o

f th

e Rí

a de

Vig

o (s

tn R

00) f

or th

e th

ree

mai

n m

odes

obt

aine

d in

EO

F an

alys

is.

Page 104: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

86

third period of vigorous northerly winds, with an average wind speed of –

10 m s–1. Finally, during September 2002, three periods can also be

defined: a first period, when southeasterly winds decreased gradually

from 6 to 3 m s–1, a second period of transition from southerly to northerly

winds; and a third period, when moderate northeasterly winds of less

than –4 m s–1 blew over the shelf (Figure 5a).

3.2. Residual currents

The low–pass filtered subtidal fluctuations of the current are presented

in Figures 2b, 6b, 8b and 10b. The current vectors are projected along the

preferred direction of water displacement: perpendicular to the transverse

section between Cabo de Mar and Punta Borneira, as obtained from

dispersion graphs. Positive values indicate inflow, whereas negative

values indicate outflow.

A 2–layered estuarine circulation pattern, characteristic of partially

mixed estuaries, was found during February 2002 (Figure 2b). The

residual circulation was positive under the influence of northerly winds

blowing on the shelf. When shelf winds changed to southerly, the

circulation pattern reversed: an inflow was recorded through the surface

layer and an outflow through the bottom layer. The level of no motion

(LNM) separating the two layers was found between 11 and 13 m during

this winter period. The cross correlation coefficient between the residual

current and the north component of shelf winds as a function of the lag

time between them is presented in Figure 6a for the 5 ADCP layers.

Page 105: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

87

a

b

-12

-6

0

6

12

15/2 18/2 21/2 24/2 27/2 2/3 5/3 8/3 11/3

m/s

15/2 18/2 21/2 24/2 27/2 2/3 5/3 8/3 11/3

10

20

30

-15-10-5051015 cm/s

Dep

th (m

)

Figure 2. (a) Low pass filtered wind velocity from the meteorological station off Cabo Silleiro during the winter period. (b) Time evolution of the low frequency component (30 h cut–off period) of the current velocity measured at stn R00 along the direction perpendicular to the section between Cabo de Mar and Punta Borneira. The dotted lines indicate the hydrographic sampling dates. Positive values of V indicate an inflow, and negative values of V indicate an outflow. The solid black line represents the level of no motion, LNM (V = 0).

Page 106: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

88

-12

-6

0

6

12

15/4 17/4 19/4 21/4 23/4 25/4

m/s

15/4 18/4 21/4 25/4

10

20

30

Dep

th (m

)

-15-10-5051015cm/s

a

b

Figure 3. (a) Low pass filtered wind velocity from the meteorological station off Cabo Silleiro during the spring period. (b) Time evolution of the low frequency component (30 h cut–off period) of the current velocity measured at stn R00 along the direction perpendicular to the section between Cabo de Mar and Punta Borneira. The dotted lines indicate the hydrographic sampling dates. Positive values of V indicate an inflow, and negative values of V indicate an outflow. The solid black line represents the level of no motion, LNM (V = 0).

Page 107: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

89

-15-10-5051015

11/7 15/7 19/7 23/7 27/7

10

20

30

Dep

th (m

)

cm/s

-12

-6

0

6

12

11/7 15/7 19/7 23/7 27/7

m/s

a

b

Figure 4. (a) Low pass filtered wind velocity from the meteorological station off Cabo Silleiro during the summer period. (b) Time evolution of the low frequency component (30 h cut–off period) of the current velocity measured at stn R00 along the direction perpendicular to the section between Cabo de Mar and Punta Borneira. The dotted lines indicate the hydrographic sampling dates. Positive values of V indicate an inflow, and negative values of V indicate an outflow. The solid black line represents the level of no motion, LNM (V = 0).

Page 108: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

90

10

20

30

Dep

th (m

)

17/9 20/9 23/9 26/9

-12

-6

0

6

12

17/9 20/9 23/9 26/9

m/s

a

b-15-10-5051015

cm/s Figure 5. (a) Low pass filtered wind velocity from the meteorological station off Cabo Silleiro during the autumn period. (b) Time evolution of the low frequency component (30 h cut–off period) of the current velocity measured at stn R00 along the direction perpendicular to the section between Cabo de Mar and Punta Borneira. The dotted lines indicate the hydrographic sampling dates. Positive values of V indicate an inflow, and negative values of V indicate an outflow. The solid black line represents the level of no motion, LNM (V = 0).

Page 109: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

91

-2.0 -1.5 -1.0 -0.5 0.0-1.0

-0.5

0.0

0.5

1.0

R(τ)

(τ, days)

L1

L2

L3 L5, L4

-2.0 -1.5 -1.0 -0.5 0.0-1.0

-0.5

0.0

0.5

1.0

R(τ)

(τ days)

L3 L4 L5

L1 L2

a b

c d

-2.0 -1.5 -1.0 -0.5 0.0-1.0

-0.5

0.0

0.5

1.0

R(τ)

(τ, days)

L1

L5 L4 L2 L3

L1

L5

L3

L2 L4

-2.0 -1.5 -1.0 -0.5 0.0-1.0

-0.5

0.0

0.5

1.0R(τ)

(τ, days)

The curve shape )(τR vs τ presented the maximum correlation, near

0.9, with a 0–0.1 d lag time for all layers, indicating an immediate response

of the ría to the remote atmospheric forcing. Layer 1 (surface) was

positively correlated with Wy. Therefore, northerly winds (Wy < 0)

originated a surface outflow (V < 0). On the contrary, layers 2 to 5 showed

a negative correlation with shelf winds.

The characteristic 2–layered circulation pattern was observed again in

the low–pass filtered transversal current velocity profile at stn R00 during

April 2002 (Figure 3b). The LNM was at 10–11 m depth, except at

beginning of the study period when the surface layer was thinner.

Specially remarkable was the transitional 3–layered circulation pattern

Figure 6. Cross correlation coefficients between the residual current in the middle Ría de Vigo (stn R00) and the North component of shelf winds recorded at the Cabo Silleiro buoy as a function of the lag time between residual currents and winds for (a) winter, (b) spring, (c) summer and (d) autumn.

Page 110: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

92

observed on April 19, with outflow through the surface and bottom layers

and inflow through the mid water column. This particular residual

current distribution was found after a wind relaxation event. The water

column responded to weak winds with weak residual currents, reaching a

maximum outflow velocity of –8 cm s–1 and a maximum inflow of 5 cm s–1.

The scalar cross correlation coefficient between residual current and shelf

winds as a function of the lag time (Figure 6b) presented a maximum

between 0 and 0.5 days for layers 1 and 2, and between 1.5 and 2 days for

layers 3, 4 and 5. The coefficient was lower than 0.5 for layers 4 and 5,

which contained the LNM when the 3–layered distribution pattern arose.

Again, the coefficient was positive for layer 1 and negative for layers 2 to

5. The relative weakness of shelf winds during the spring period was the

reason behind the poor correlation coefficients.

The low–pass filtered transversal current velocity profile at stn R00

during July 2002 (Figure 4b) was characterised by a strong surface outflow

reaching –12 cm s–1 during the first period of upwelling–favourable

northerly winds. The second period corresponded to a profound

relaxation of shelf winds, which produced a reversal of the residual

circulation pattern; the surface inflow extended over the upper 12 m with

a celerity of <7 cm s–1 and the bottom outflow was lower than –3 cm s–1.

Finally, intense shelf northerly winds changed again the circulation

pattern during the third period, with surface and bottom currents similar

to the first period. The cross correlation coefficient (Figure 6c) between the

residual current and the north component of shelf winds was about –0.8

for layers 2 to 5, whereas for layer 1 was just >0.6, because it included the

LNM. The curve shape indicates that the correlation was independent of

lag times ranging from 0 to –48 h.

Page 111: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

93

During September 2002, the residual current pattern was initially

negative in response to the dominant southerly winds blowing over the

shelf. Maximum surface inflow currents of 10 cm s–1 and bottom outflow

currents of –10 cm s–1 were recorded (Figure 5b). The LNM was deeper (17

m) than in the previous downwelling episodes of February and July 2002.

In response to the relaxation of southerly winds, the reversal of the

residual pattern persisted but current velocities diminished and the LNM

deepened. During the transition from southerly to northerly shelf winds,

on September 23, the circulation evolved to a 3–layered pattern, with

outflow through the surface and bottom layers and inflow through the

intermediate layer. The subsequent dominance of northerly winds re–

established the 2–layered positive circulation pattern, typical of upwelling

favourable winds, with an average velocity of about –5 cm s–1 in the

surface layer and 5 cm s–1 in the bottom layer. The LNM was shallower

(~10 m) than during the previous downwelling conditions. Finally,

although shelf winds change again to southerly, the intensity was not

enough to produce a reversal of the residual circulation pattern. The cross

correlation coefficient (Figure 6d) between residual currents and shelf

winds was near 0.9 for all layers except 2 and 3, which contained the

LNM. All layers exhibited the maximum correlation near 0.5 days and the

correlation diminished abruptly for lag times larger than 1 day. The

correlation was positive for layers 1 and 2, and opposite for the rest

indicating that southerly winds (Wy > 0) favoured the entry of water

through the upper layers 1 and 2, while the outflow was located in layers

3 to 5.

Page 112: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

94

3.3. Thermohaline variability

A CTD probe was dipped at five hydrographic stations (R00 to R04)

along the transect from Cabo de Mar to Punta Borneira (Figure 1).

Completion of the transect took about 20 minutes. It was repeated four

dates in February (Figure 7), April (Figure 8), July (Figure 9) and

September (Figure 10) with a periodicity of ½ wk.

During February 2002, a brief thermal inversion caused by winter

cooling was observed. Heat exchange with the atmosphere (H) was

negative (Figure 11a), although the loss of energy diminished from the

beginning to the end of the study period. The hydrological balance ranged

from 5 to 20 m3 s–1. The dominant northerly winds at the beginning of the

study period caused the entry of very salty (>35.8) ENACW throughout

the bottom layer of the ría (Figure 7a), that upwelled to the surface layer.

Upwelling of ENACW produced a marked increase of salinity,

specially in the northern side of the transverse section (Figure 7b). In

response to the subsequent relaxation of shelf northerly winds, ENACW

pulled back to the shelf and the bottom salinity decreased to 35.6 (Figure

7c). Finally, the dominant southerly winds from February 25 on,

produced a reversal of the residual circulation pattern, which piled the

fresh water runoff in the surface layer (S <35.2) and evacuated the bottom

waters of the ría (S = 35.4) through the bottom outgoing current (Figure

7d). Since the thermal inversion maintained during the whole period

(Figure 5e–h), salinity controlled the hydrographic variability and the

gravitational circulation.

Page 113: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

95

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

18/02/02 18/02/02

21/02/02 21/02/02

25/02/02 25/02/02

28/02/02 28/02/02

S N S N

34 34.3 34.6 34.9 35.2 35.5 35.8 11.0 12.8 14.6 16.4 18.2 20.0

R01 R03R00R04 R02 R01 R03R00R04 R02

a

b

c

d

e

f

g

h

Figure 7. Contour plots of salinity (a–d) and temperature (e–h) along the section between Cabo de Mar and Punta Borneira from 11th to 22th April 2002.

Page 114: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

96

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

11/04/02 11/04/02

15/04/02 15/04/02

18/04/02 18/04/02

22/04/02 22/04/02

S N S N

34 34.3 34.6 34.9 35.2 35.5 35.8 11.0 12.8 14.6 16.4 18.2 20.0

R01 R03R00R04 R02 R01 R03R00R04 R02

a

b

c

d

e

f

g

h

Figure 8. Contour plots of salinity (a–d) and temperature (e–h) along the section between Cabo de Mar and Punta Borneira from 11th to 22th April 2002.

Page 115: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

97

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

15/07/02 15/07/02

18/07/02 18/07/02

22/07/02 22/07/02

26/07/02 26/07/02

S N S N

34 34.3 34.6 34.9 35.2 35.5 35.8 11.0 12.8 14.6 16.4 18.2 20.0

R01 R03R00R04 R02 R01 R03R00R04 R02

a

b

c

d

e

f

g

h

Figure 9. Contour plots of salinity (a–d) and temperature (e–h) along the section between Cabo de Mar and Punta Borneira from 15th to 26th July 2002.

Page 116: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

98

-40

-30

-20

-10

0

S N

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

1000 1500 2000

-40

-30

-20

-10

0

17/09/02 17/09/02

19/09/02 19/09/02

23/09/02 23/09/02

26/09/02 26/09/02

S NR01 R03R00R04 R02 R01 R03R00R04 R02

34 34.3 34.6 34.9 35.2 35.5 35.8 11.0 12.8 14.6 16.4 18.2 20.0

a

b

c

d

e

f

g

h

Figure 10. Contour plots of salinity (a–d) and temperature (e–h) along the section between Cabo de Mar and Punta Borneira from 17th to 26th September 2002.

Page 117: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

99

5-9 10-9 15-9 20-9 25-9 30-9-3

0

3

6

9

12

15 Q P E H

day

Q -

P - E

(m3 /s

)

-200

-150

-100

-50

0

50

H (calcm

-2día-1)

5-4 10-4 15-4 20-4 25-4 30-4

0

10

20

30

40

50

Qr P E H

day

Q -

P - E

(m3 /s

)

-100

-50

0

50

100

150

200

250

H (calcm

-2día-1)

5-2 10-2 15-2 20-2 25-2

0

20

40

60

80

100 Qr P E H

day

Q -

P - E

(m3 /s

)-250

-200

-150

-100

-50

H (calcm

-2día-1)

5-7 10-7 15-7 20-7 25-7 30-7-3

0

3

6

9

12

15 Qr P E H

day

Q -

P - E

(m3 /s

)

0

100

200

300

400H

(calcm-2día

-1)

a b

c d

Figure 11. Hydrological balance terms: Q (continental runoff), P (precipitation) and E (evaporation) in m3 s–1 (left axis), and radiative balance, H, in cal cm–2 d–1 (right axis) for the (a) winter, (b) spring, (c) summer and (d) autumn periods. The shaded areas indicates the sampling periods.

During April 2002, a marked haline stratification produced by

continental runoff waters occurred at the beginning of the study period

(Figure 8a). The halocline was situated at 10–15 m. On the contrary,

temperature was homogeneously distributed, ranging from 13.5 to 13.7ºC

(Figure 8e). The thermal inversion observed in February 2002 had

practically disappeared. Reduction of the fresh water flow (Figure 11b)

produced a weakening of the haline stratification (Figure 8b). Surface

salinity was not homogenous: lower values were recorded in the

Northern side of the transect because of the Coriolis deflection to the right

of continental runoff. At the bottom, salinity increased due to the entry of

ENACW in response to the dominant northerly winds. A marked change

in the distribution of temperature was observed on April 18 (Figure 11g);

Page 118: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

100

it can be observed how a weak surface thermocline was established (near

20 m depth) due to the differences between the colder oceanic water that

was introduced into the ría through the bottom layers and the surface

warmer fresh water. Finally, the increasing solar irradiation (Figure 11b)

was able to produce a marked thermal stratification at the end of the

study period. These are the appropriate conditions for the development of

the spring bloom in the rías (Nogueira el al. 1997)

During July 2002, temperature rather than salinity gradients controlled

the density distribution in the middle ría (Figure 9). Continental runoff

was very low throughout the study period (< 9 m3 s–1; Figure 11c). On July

15, the salinity and temperature distributions showed the typical response

of the water column to an upwelling event: the northern side of the

transect was fresher (Figure 9a) and warmer (Figure 9e) than the southern

side. The thermocline was at about 10 m depth. Despite the subsequent

relaxation of shelf winds, the effect of coastal upwelling still remained in

the water column 3 days later, when higher salinities (Figure 9b) and

lower temperatures (Figure 9f) were recorded. The reversal of the

circulation in response to the shelf winds calm (Figure 4) had a clear

impact on the water column: salinity was more homogenous and the

halocline was difficult to distinguish (Figure 9c) while surface

temperature increased and the maximum gradient was found over 20 m

depth (Figure 9g). Shelf water tended to enter the ría through the surface

layer and forced to sink. This downwelling process was favoured because

river discharge was negligible. From July 23 onwards, intense and

persistent northerly winds changed again the circulation pattern. The

reduced river discharge left the estuary by the surface layers and saltier

Page 119: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

101

(Figure 9d) and colder water (Figure 9h) entered into the ría through the

bottom layer.

Finally, during September 2002, the freshwater flow was again very

limited: from 5.7 to 14.5 m3 s–1 (Figure 4d). The thermohaline structure

found on September 17 corresponded with a typical autumn downwelling,

with warm (>17ºC) and relatively salty (>34.8) shelf surface waters

occupying the surface layer of the middle ría (Figure 10a, e). The

persistence of shelf southerly winds produced a deepening of the surface

salinity and temperature as a consequence of downwelling; bottom

salinities were as low as 35.0 (Figure 10b, c) and bottom temperatures

increased to 16ºC (Figure 10f, g) from September 17 to 23. Downwelling

contributed to homogenise the salinity through the water column, but the

solar irradiation was able to keep the temperature gradient. Southerly

winds changed to northeasterly on September 22 and this effect was

recorded in the water column with a slight uplift of the isohalines (Figure

10d) and isotherms (Figure 10h), suggesting the entry of oceanic ENACW

through the bottom layers on September 26.

4. Discussion

The sampling programme allowed to capture the hydrographic

(salinity and temperature distributions) and dynamic (residual circulation

profiles) response of the water column over the wide range of 1) remote

and local winds intensity and direction and 2) thermal and/or haline

stratification/mixing conditions in a coastal upwelling embayment of the

NW Iberian coast. The winter period (February 2002) was characterised by

a marked thermal inversion under an unexpectedly low continental

runoff, as compared with a long term average (Nogueira et al. 1997).

Winds were very variable in magnitude and direction during the incipient

Page 120: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

102

stratification of the spring period (April 2002). During the summer period

(July 2002), a succession of short–time upwelling events separated by

intervals of stratification was recorded, a pattern typical of the upwelling

season of coastal upwelling systems at temperate latitudes (Alvarez–

Salgado et al. 1993; Hill et al. 1998). Finally, during the autumn

(September 2002), the pattern reversed and short–time downwelling

events, separated by intervals of stratification were observed.

4.1. Remote winds, a key forcing agent of the subtidal motion in

estuaries and coastal inlets

Since the middle 1970’s the significance of shelf wind forcing in driving

the motion in estuaries and bays has been recognised (Carter et al. 1979).

In the last two decades, the effort concentrated on the wind–induced

subtidal variability in partially mixed estuaries of diverse sizes, such as

the San Francisco Bay (Walters 1982; Walters and Gartner 1985), the

Delaware Bay (Wong and Garvine 1984; Wong and Mosses–Hall 1998)

and the Chesapeake Bay (Vieira 1985, 1986; Valle–Levinson 1995; Valle–

Levinson et al. 1998). Furthermore, atmospheric forcing is also relevant for

the subtidal variability in coastal lagoons with restricted communication

with the ocean (e.g. Wong and Wilson 1984; Wong and DiLorenzo 1988).

These studies revealed that the subtidal variability in estuaries is mainly

induced by winds, through a combination of remote and local effects.

The effect of shelf winds is particularly relevant in an embayment

affected by coastal upwelling. This is the case of the Rías Baixas in North

West Spain (Rosón et al. 1997; Alvarez–Salgado et al. 2000; Pardo et al.

2001; Gilcoto et al. 2001), Asysen Fjord in Southern Chile (Cáceres et al.

2002) or Bantry Bay in South West Ireland (Edwards et al. 1996).

Page 121: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

103

All these coastal inlets are located in temperate latitudes, where

upwelling is not permanent but seasonal and repeated wind

stress/relaxation cycles of period 1–2 wk occur during the upwelling

season (Hill et al. 1998). For the case of the NW Iberian upwelling, a

sampling strategy consisting on visiting the study site every 3–4 d was

commonly applied during the 1990’s to asses the response of the water

column to the shelf wind stress/relaxation cycles (Rosón et al. 1997;

Álvarez– Salgado et al. 2000; Pardo et al. 2001). These authors stated that,

at the time scale of the sampling frequency, the water column responded

immediately to shelf wind stress. Later, Gilcoto et al. (2001) observed that

the subtidal circulation of the surface layer of the Ría de Vigo was delayed

less than ½ wk compared with shelf winds, after analysing the data

obtained with a mechanic current meter deployed during a short period in

September 1990. Therefore, they demonstrated that a sampling frequency

of 3–4 days was insufficient to study the lag time between subtidal

circulation and shelf winds. Combination of continuous records of shelf

winds and residual currents recorded at 10 min intervals in the present

work allowed to quantify this lag time by means of a cross correlation

analyses. Lag times from 0 to 24 h at 6 h intervals were tested. Maximum

correlation coefficients (>65%) between shelf winds and residual current

records were obtained for lag times of 12 hours in February and

September. In July, the correlation coefficient was not different for shelf

winds blowing over the last 24 hours (Table 2). The short lag times

observed in the Ría de Vigo are probably related to the relatively small

size of the rías compared with other embayments. In this sense, the small

size of San Francisco Bay was the reason argued by Walters (1982) to

explain the higher frequency response in comparison with the much

Page 122: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

104

larger Chesapeake Bay. Wong (2002) found that remote wind–induced

coastal pumping effect is by far the most important mechanism to force

the subtidal current fluctuations in Chesapeake Bay over time scales of 2–

3 d. The percentage of the observed subtidal current variability explained

by shelf winds in the Ría de Vigo (>65%) was much larger than the 25%

found in the Potomac Estuary by Elliot (1978) and of the same order of the

70% found in the Chesapeake Bay by Wong (2002).

A model II linear regression analysis (Sokal et al. 1994) was applied to the

offshore Ekman transport (m3 s–1 km–1) calculated from shelf winds and

the surface fluxes (m3 s–1) recorded at stn R00 with the ADCP current

meter. The results are presented in Table 2 for the winter, summer and

autumn periods. The slopes of the regression lines are comparable with

the wideness of the surface layer of the middle Ría de Vigo (2.2 km). This

observation indicates that the middle ría is under the direct influence of

shelf wind stress. A similar result was obtained by Rosón et al. (1997) in

the Ría de Arousa and Álvarez–Salgado et al. (2000) in the Ría de Vigo,

with their inverse method. It should be noted again that the results of an

inverse method are applicable to the time scale of a sampling frequency

(½ wk) that largely exceed the lag time of response of the rías to shelf

winds. However, the results of the direct and continuous measurements of

this work indicate that the ría behaves as an extension of the shelf at the

short time scale of the lag time.

Page 123: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

105

February Time lag (h) Slope (km) y–intercept (m3 s–1) R R2

0 –2.3 ± 0.1 497±72 –0.89 0.79 –6 –2.4 ± 0.1 552±75 –0.89 0.79

–12 –2.4 ± 0.1 594±91 –0.84 0.71 –18 –2.4 ± 0.1 624±112 –0.77 0.59 –24 –2.4 ± 0.2 640±133 –0.67 0.45

average (0, –12) –2.4 ± 0.1 569±75 –0.89 0.79 July

Time lag (h) Slope (km) y–intercept (m3 s–1) R R2 0 –2.1 ± 0.1 1386 ± 124 –0.75 0.56 –6 –2.0 ± 0.1 1323 ± 111 –0.78 0.61

–12 –2.0 ± 0.1 1766 ± 119 –0.80 0.64 –18 –2.0 ± 0.1 1709 ± 116 –0.80 0.64 –24 –2.1 ± 0.1 1158 ± 116 –0.80 0.64

average (0, –12) –2.1 ± 0.1 1360 ± 11 –0.79 0.63 average (0, –18) –2.1 ± 0.1 1353 ± 106 –0.81 0.65

September Time lag (h) Slope (km) y–intercept (m3 s–1) R R2

0 –2.6 ± 0.2 –233 ± 92 –0.75 0.56 –6 –2.6 ± 0.2 –261 ± 84 –0.80 0.63

–12 –2.5 ± 0.3 –296 ± 82 –0.80 0.63 –18 –2.4 ± 0.3 –332 ± 88 –0.77 0.59 –24 –2.3 ± 0.3 –369 ± 89 –0.77 0.59

average (0, –12) –2.6 ± 0.2 –308 ± 80 –0.81 0.65 Table 2. Analysis of the regression between the Offshore Ekman transport calculated from shelf winds (Cabo Silleiro) and the surface flow calculated with the DCM12 current meter moored at stn R00 during February, July and September for time lags of 0 to –24 h at –6 h intervals. The 95% confidence interval of the slope and y–intercept are provided.

The y–intercept of the regression equations in Table 2 represents the

intrinsic circulation of the ría after removal of the variability due to shelf

winds. A positive y–intercept indicates a negative intrinsic circulation

(surface inflow) and a negative value indicates a positive intrinsic

circulation (surface outflow). In view of these results, a negative intrinsic

Page 124: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

106

circulation was found in July and February, while in September it was

positive.

Although the 2–layered residual circulation scheme is the most

common in the Ría de Vigo, a 3–layered residual circulation develops

under the influence of low–intensity shelf winds during the transition

from upwelling to downwelling favourable conditions and vice versa. It

consists of an outflow through the surface and bottom and an inflow

through the intermediate layer. Álvarez–Salgado et al. (1998) detected the

same pattern during a 1 day period in September 1991 in the Ría de Vigo.

This pattern has also been found by Valle–Levinson et al. (2002) studying

the effect of winds on the exchange flow of a channel constriction of the

Chilean Inland Sea. They found a 3–layer response to winds that opposed

the density–induced near surface outflow and corresponded with other

observations (Svendsen et al. 1978; Cáceres et al. 2002) and numerical

simulations (Klinck et al. 1981) in other systems. The circulation pattern of

an archetypal fjord also consists of 3 layers: a surface layer, flowing

seawards, driven by river discharge at the head of the fjord; an

intermediate layer, flowing landwards; and a deep layer, which is only

replaced when the intermediate inflow is sufficiently dense to displace the

existing deep water. Although the morphology and origin of the rías are

different from fjords, they have analogous characteristics. In fact, wind

forcing can be the main driving mechanism in both systems and the two

systems present evidences of wind–induced upwelling; Aysen Fjord, on

the southern coast of Chile (Cáceres et al. 2002), presented a 3–layer

structure that was consistent with up–fjord wind–induced exchange. The

3–layered circulation could be understood as a transitional adjustment of

Page 125: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

107

the system to the dominant 2–layered circulation commonly found in the

Ría de Vigo.

4.2. Coupling between hydrography and hydrodynamics

In the present work, residual flows in the middle Ría de Vigo have also

been assessed from the short–time scale variability of the distributions of

salinity and temperature by means of the inverse method described in

section 2.3. Simultaneous ADCP measurements allowed comparison of

two independent methods for calculating water flows. Figure 12 presents

the relationship between measured (ADCP) and calculated (inverse

method) water flows. The 1:1 line explained 94% of the variability of the

calculated flows. Therefore, both methods corroborate the validity of the

calculated water fluxes. Gilcoto et al. (2001) obtained the first comparison

between measured and calculated flows, matching the output of an

inverse method and the surface residual currents obtained with a

mechanic current meter deployed at 3 m depth in the central Ría de Vigo.

However, Gilcoto et al. (2001) data reduce to just four observations during

a 2 wk period in September 1990.

Water displacements evaluated from direct current measurements are

restricted to the mooring position, a single depth in the case of the

mechanic current meter of Gilcoto et al. (2001), while the inverse method

extended to the complete cross section. Furthermore, the possibility of

splitting the residual fluxes calculated with the inverse method into 3

terms allows to asses their relative contribution to the total residual flow.

Page 126: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

108

-1200 -800 -400 0 400 800 1200 1600-1200

-800

-400

0

400

800

1200

1600Q

B(adc

p) (m

3 /s)

QB(box model) (m3/s)

Figure 12. Linear regression between bottom water fluxes calculated with the inverse box model of Álvarez–Salgado et al. (2000) and the measurements preformed with the DCM12 current meter.

Table 3 shows 1) the time evolution of the bottom water flux in the middle

ría, BQ , calculated with the inverse method; 2) the non–steady–state term

of BQ , ( )NSSBQ ; and 3) the steady–state term of BQ , divided in a

hydrological term, HSSBQ ))(( , and a radiative term, RSSBQ ))(( . An

analysis of the variance of the 3 terms (σNSS2, σH2 and σR2) yielded that the

short–time scale evolution of BQ was primarily controlled by the non–

steady–state term (72% of σNSS2+σH2+ σR2), then by the radiative term (24%)

and, residually by the hydrological term (4%).

The non–steady state term, which depended on the temporal changes

in salinity ( tS ∆∆ / ) and temperature ( tT ∆∆ / ), correlated significantly

with the offshore Ekman transport, – XQ , considering the 16 pairs of data

obtained by Álvarez–Salgado et al. (2000) combined with the 12 pairs

obtained in this work:

Page 127: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

109

( ) XNSSB Q2091130612Q ).(.)( ±+±−=

R = 0.83, n = 28, p < 0.0001 (23)

Period Days (QB)R (QB)H (QB)NSS QB (m3/s) (m3/s) (m3/s) (m3/s) 18/21 572 136 346 1054

February 21/25 599 77 –608 67 25/28 94 2533 –3257 –630 11/15 471 –1 1496 1967

April 15/18 442 108 260 811 18/22 199 300 –240 258 15/18 61 628 526 1215

July 18/22 24 436 –819 –358 22/26 16 442 550 1009 17/19 691 71 –1614 –851

September 19/23 1009 40 –790 259 23/26 274 –183 471 562

Table 3. Time evolution of the bottom flow in the middle segment of the Ría de Vigo, QB, between two consecutive surveys during the four study periods as calculated with the inverse method of Álvarez–Salgado et al. (2000). The hydrological, (QB)H , radiative, (QB)R, and non–steady–state term, (QB)NSS are separated.

In view of these results, about 70% of the variability in the bottom flow,

at the time scale of the sampling frequency (1/2 wk), depended on the

non–steady state term and the offshore Ekman transport controlled 70% of

the variability of the non–steady–state term.

In addition, combing the data of surface water flows, SQ , continental

runoff, RQ , and offshore Ekman transport, – XQ , obtained by Álvarez–

Salgado et al. (2000) from 1990 to 1997, with the terms obtained in this

work, produced the following empirical equation:

Page 128: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

110

X3

R3

S Q102022Q10413Q −− ±−±= ).(.)(

R = 0.92, n = 27, p < 0.001 (24)

The regression coefficients of eq. (24) are not significantly different

from those obtained by Álvarez–Salgado et al. (2000), indicating that the

2002 data were consistent with the previous fit. Note that the coefficient of

–QX, 2.2±0.2 km, is again comparable with the wideness of the surface

layer of the middle Ría de Vigo.

A second way to compare the results obtained from direct

measurements and inverse method calculations is the depth of the level of

no motion (LNM) between the outgoing (ingoing) surface current and the

outgoing (ingoing) bottom current obtained with both methods.

Traditionally the LNM in box models was set at the depth of the halocline

(Pritchard 1952; Officer 1980). In the case of the Rías Baixas, since both the

salinity and temperature profiles are considered, the LNM has been

preferentially set at the pynocline (Rosón et al. 1997; Álvarez–Salgado et al.

2000; Gilcoto et al. 2001; Pardo et al. 2001). In this work, this hypothesis

has been corroborated: the difference between the two sets of LNM depths

was 0 ± 2.5 m.

A significant transversal variability was found in the salinity and

temperature distributions of the middle segment of the Ría de Vigo. In

general, the northern margin was fresher and warmer than the southern

margin of the ría. This gradient is caused by the Coriolis deflection of

continental waters under a positive residual circulation pattern. The same

reason is applicable to the entry of the cold and salty ENACW through the

southern bottom margin. In addition, the 3–D residual circulation pattern

of the outer ría (Souto et al. 2003) seems to play a significant role in the

Page 129: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

111

asymmetry of the residual currents in the middle ría (Gilcoto 2004) and,

therefore in the transversal variability of the thermohaline properties.

Finally, an additional source of variability of the thermohaline

structure of the middle ría is the quality of the ENACW upwelled over the

NW Iberian shelf that enters the ría: it was warmer and saltier in winter

(>35.8) than in summer (<35.6). The Galician Rías Baixas are located at the

boundary between the temperate and subpolar regimes of the Eastern

North Atlantic, where two branches of ENACW can upwell. These two

branches were described first by Fraga et al. (1982), who observed that

they meet in a quasi permanent subsurface front that migrates seasonally

from south of River Miño in winter to east of River Eo in summer (Figure

1) according to Castro (1997). Ríos et al. (1992a) characterized the origin of

both branches: 1) ENACW of subpolar origin (ENACWp) represents the

water bodies formed to the north of Cabo Fisterra; and 2) ENACW of

subtropical origin (ENACWt) formed to the south of Cape Roca. These

authors established that ENACWt has a salinity >35.66 and ENACWp

<35.66. During the downwelling season, warm and saltier surface waters

of subtropical origin are promoted by the Iberian Poleward Current off

the Rías Baixas (Haynes and Barton 1990, 1991). They pile on the shelf and

enter the rías through the surface or bottom layer depending on shelf

winds. Therefore, the origin of upwelled water into the rías depends on: 1)

the seasonal migration of the front and 2) the vertical displacements of

ENACW promoted by the shelf winds: northerly winds of high intensity

produce the entry of the deeper ENACWp branches into the ría during the

summer. A similar subsurface front between water masses is found off

Cape Blanco, NW Africa (21º N). South of the cape, SACW (South Atlantic

Central Water) upwells, with lower salinity and higher nutrient

Page 130: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

112

concentrations than the NACW (North Atlantic Central Water) that

upwells north of the cape (Fraga et al. 1985).

Page 131: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Thermohaline variability and residual circulation in the Ría de Vigo

113

5. Conclusions

Although the basic knowledge on the hydrodynamics of the Spanish

Rías Baixas developed over the last decade with inverse models based on

hydrographic data, direct current measurements are still scarce, with a

poor spatial and temporal coverage. This study was specifically designed

to analyse the short time and space scale variability of the hydrodynamics

of the Ría de Vigo, the paradigm of a coastal embayment under the

influence of coastal upwelling, from an intensive recording programme of

the vertical structure and displacements of the water column.

The following conclusions can be drawn:

1. Remote shelf winds controlled more than 65% of the variability

observed in the subtidal circulation of the ría, whereas the heat exchange

with atmosphere represented less than 25% and continental runoff less

than 5% at an annual basis. Local winds did not contributed significantly

to the subtidal circulation of the ría.

2. The subtidal circulation of the ría responded to the effect of shelf

wind with an average lag time ranging from 0 to 2 days, depending on the

layer and the time of the year. Commonly, the lag is less than 1 day. This

rapid response compared with other estuaries and coastal embayments is

probably related to the comparatively small size of the ría and the V–

shaped topography.

3. The 2–layered residual circulation pattern, assumed with the inverse

models and calculated with the numeric models of the Ría de Vigo, has

been confirmed by repetitive measurements in the central part of the

embayment under contrasting hydrographic and meteorological

conditions. In addition, a 3–layered circulation pattern has been identified

during the transition from positive to negative circulation and vice versa.

Page 132: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 2

114

Page 133: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3:

Lateral variability in the coastal upwelling system

of the Ría de Vigo

Page 134: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 135: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

117

Chapter 3, the research work presented in this chapter is also a

contribution to the paper:

S. Piedracoba, J. L. Herrera, G. Rosón, M. Gilcoto, C. Souto, X. A.

Álvarez–Salgado. Lateral variability in the coastal upwelling system of

the Ría de Vigo (NW Spain). J. Mar. Sys. (submitted).

Abstract: Ship mounted and moored ADCP observations and the

HAMSON numerical model have been combined to produce a realistic

view of the lateral variability of the subtidal circulation in the coastal

upwelling system of the Ría de Vigo. The lateral differences cannot be

explained considering only rotational effects (Coriolis acceleration), but

also 1) the interaction of shelf wind–driven currents with the mouths of

the ría for the case of the longitudinal component, and 2) the spatial

gradient of local wind for the case of the transversal component of the

current. Both remote and local winds could be responsible for non–linear

advection across the estuary, i.e. other accelerations (advection and

friction terms) that play an important role in the transverse dynamics of

estuaries and coastal embayments.

Page 136: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

118

Page 137: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

119

Resumen: Se obtuvieron medidas de corriente con un perfilador

acústico de efecto Doppler instalado a bordo del B/O Mytilus y, con el

fondeo de un correntímetro doppler en el centro de la ría de Vigo. Estas

medidas se combinaron con resultados los del modelo numérico

HAMSON para dar una descripción realista de la variabilidad lateral de la

circulación residual en el sistema de afloramiento costero de la Ría de

Vigo. Las diferencias laterales en las componentes longitudinal y

transversal de la corriente no pueden explicarse atendiendo únicamente a

efectos rotacionales (aceleración de Coriolis), sino también a 1) la

interacción de las corrientes inducidas por el viento de plataforma con las

bocas de la ría para el caso de la componente longitudinal, y a 2) el

gradiente espacial del viento local para el caso de la componente

transversal. Tanto el viento costero como el local pueden causar advección

a lo largo de la ría, es decir, otras aceleraciones no lineales (términos de

advección y fricción) que juegan un papel importante en la dinámica de

estuarios y otros sistemas costeros.

Page 138: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

120

Page 139: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

121

1. Introduction

Primary factors controlling the physics of estuaries are tidal motion,

atmospheric forcing, and freshwater discharge (Reynolds–Fleming et al.

2004). The influence of atmospheric forcing on the subtidal variability has

been demonstrated in a variety of coastal systems (Wang and Elliot 1978;

Elliot et al. 1978; Wong and Garvine 1984). Wind driven variability occurs

mainly due to remote wind effects, local winds or a combination of both.

Remote wind effects come predominantly from along–shore winds, which

produce coastal sea level fluctuations along the shelf and at the mouth of

the estuary owing to cross–shelf Ekman transport (Janzen and Wong

2002). Local wind effects induce also estuarine sea level and current

fluctuations. Other factors can modify the response to remote and local

winds. These include bathymetry (Csanady 1973; Hunter and Hearn 1987;

Signell et al. 1990; Wong 1994), size and orientation of the estuary

(Garvine 1985) and lateral dimensions and stratification, which can

modify the importance of the Coriolis effect (Asplin 1995).

Although subtidal motion is inherently a 3–D process, most studies on

estuarine dynamics have focused primarily on the axial variability.

Increased attention has been paid to the lateral structure of the density

and along–estuary mean flow fields (e.g. Kjerfve 1978; Kjerfve and Proehl

1979; Huzzey and Brubaker 1988; Wong 1994; Wong and Münchow 1995;

Valle–Levinson and Lwiza 1995). On the other hand, the lateral variability

of the flow field, which provides the complete 3–D description of the

estuary, has received less attention. There is some evidence, however, that

it may have important repercussions for the residual volume transport

and salt balance in estuaries and coastal embayments (Dyer 1974).

Therefore, the study of a hydrodynamic system requires resolution of the

Page 140: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

122

longitudinal and transversal components of the momentum balance. The

lateral dynamics can be sometimes approximated to geostrophic (Dyer

1997). However, this balance can be modified by non–linearities arising

from different sources that make the friction and advection accelerations

relevant to the transverse dynamics. Thus, secondary flows or lateral

accelerations arise from a combination of several interconnected sources

(Dyer 1997; Cáceres et al. 2004) such as friction, topography or bottom

profiles (the cross section of the estuary is not rectangular; the vertical

exchange between water of different density is not homogenously

distributed, etc). Morphology and sharp changes in the bathymetry (the

presence of sills or coastline contractions) can promote abrupt

longitudinal accelerations and deflections of the flow).

The study of time–dependent lateral variations in physical/chemical

properties (lateral distributions of inorganic nutrients, oxygen,

chlorophyll a and bacterioplankton) revealed influences on the

biogeochemistry of estuaries and coastal embayments. For example,

lateral differences in phytoplankton productivity were observed (Malone

et al. 1986) and remote sensing has been used to describe the lateral

variability of phytoplankton chlorophyll a (Weiss et al. 1997) in

Chesapeake Bay.

This is the first time that the lateral differences in the longitudinal and

transversal current pattern in the coastal upwelling system of the Ría de

Vigo (NW Spain) are studied through direct measurements of currents,

obtained from vessel mounted and moored Acoustic Doppler Current

Meter (ADCP), together with a numerical model. A description of the

main characteristics of the mean residual flows, emphasizing on their

Page 141: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

123

spatial structure under different meteorological, hydrographic and

dynamic scenarios was conducted.

2. Material and methods

2.1. Study site

The NW coast of the Iberian Peninsula is located at the boundary

between the subtropical and subpolar regimes of the Eastern North

Atlantic, where upwelling–favorable shelf northerly winds predominate

from March–April to September–October in response to the seasonal

migration of the Azores High. On the contrary, downwelling–favorable

southerly winds prevail the rest of the year (Wooster et al. 1976; Bakun

and Nelson 1991). However, this seasonal pattern explains <20% of the

variability of shelf winds, whereas >70% et al. concentrates at frequencies

<30 days (Nogueira et al. 1997). Álvarez–Salgado et al. (1993) observed

that the frequency of upwelling episodes at 42º 30’N was 14 ± 4 days.

The Ría de Vigo is one of the Rías Baixas, four large (>2.5 km3) and V–

shaped coastal inlets located in the NW Iberian Peninsula, connected with

the adjacent shelf. They behave as an extension of the shelf with a two–

layered residual circulation pattern, positive under upwelling and

negative under downwelling conditions (Piedracoba et al. 2005a,

Chapter 2).

Previous works about the subtidal circulation of the Rías Baixas

referred mainly to the longitudinal component of the current, because this

component is the responsible for the exchange between the rías and the

shelf: in the Ría de Arousa (Rosón et al. 1997), the Ría de Vigo (Álvarez–

Salgado et al. 2000; Gilcoto et al. 2001; Piedracoba et al. 2005a, Chapter 2)

and the Ría de Pontevedra (Pardo et al. 2001). Although lateral differences

have been observed before in the density distribution (Otto 1975; Rosón et

Page 142: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

124

al. 1997) and the horizontal velocity field (Gómez Gallego 1971; Castillejo

and Lavín 1982) of the Ría de Arousa, only the longitudinal circulation

scheme was studied in the Ría de Vigo. Some previous works focused on

the flows that enter/leave the Ría de Vigo using moored instruments

(Gilcoto et al. 2001; Souto et al. 2001; 2003; Piedracoba et al. 2005a,

Chapter 2), but they did not have the spatial resolution required to

elucidate lateral differences in the flow field.

Numerical simulations of the Ría de Vigo by Souto et al. (2003)

predicted the lateral variability of currents. An asymmetry between the

flows through the northern and southern mouths weakened towards the

middle and inner ría. Gilcoto (2004) also studied the lateral variability of

the longitudinal and transversal components of the flow in the Ría de

Vigo and found with a 3–D box model an asymmetry of the residual

currents in the central part of ría. More recently, significant lateral

differences were found in the salinity and temperature distributions of the

middle segment of the Ría de Vigo by Piedracoba et al. (2005a, Chapter 2).

2.2. Data collection

The measurement of net fluxes across estuaries requires acquiring

current velocity data at a number of stations distributed along the entire

bathymetric range to represent the full variation through the cross–

section. This is a source of potentially large errors. Kjerfve et al. (1981)

suggested that, to minimise the percentage of error in the fluxes to <15%

in estuaries with dimensions similar to North Inlet (34 Km2 salt marsh),

measurements should be made every 2·106 m2.Surveys were made along a

transverse section in the central Ría de Vigo (Figure 1). Current velocity,

salinity and temperature profiles were measured aboard R/V Mytilus

with an RDI Vessel Mounted Broad Band Acoustic Doppler Current

Page 143: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

125

Profiler (VMADCP) operating at 300 kHz and a Sea–Bird SBE–25

conductivity–temperature–depth (CTD) probe.

Surveys were made on 18 and 21 February, 11 and 22 April, 18 and 22 July

and 19 September 2002 to investigate contrasting meteorological and

hydrographic conditions.

9.0º 8.9º 8.8º 8.7º 8.6ºLongitude (West)

42.1º

42.2º

42.3º

Latit

ude

(Nor

th)

Rande Strait San

Sim

on B

ay

C. MarCíes Islands

1000 1500 2000

-40

-30

-20

-10

07.3-10.8NS CS

5.9-15.2SS

6.3-10.6

NB SB6.2-15.3

N(R03) S(R01)

9.5-17.0

P.Borneira

DCM12 st.CTD station

Vigo

Pontevedra

Arousa

River Oitabén Verdugo

Silleiro buoy

Rande Strait

Iberian Peninsula

Figure 1. Map of the Ría de Vigo. The transverse section between Cabo de Mar and Punta Borneira is shown together with the range of variation of the area for each sector: northern surface (NS), central surface (CS), southern surface (SS), northern bottom (NB) and southern bottom (SB), in (*103 m2). The inset shows the position of the Ría de Vigo within the Rías Baixas and the Seawatch buoy of Puertos del Estado off Cape Silleiro.

The details of each sampling are summarized in Table 1.

The transect was occupied between 25 and 31 times during each of the

seven data collection intervals. The VMADCP data were averaged over 60

s from ~12 min in the mid–channel to ~24 min at the northern and

Page 144: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

126

southern ends of the transect. At the same time, current profiles were

measured with an Aandera DCM12 Acoustic Doppler Current Meter

(ADCP) moored at the central station of the transect (Figure 1). The

mooring was deployed from 15 February to 11 March, from 11 to 25 April,

from 10 to 29 July and from 17 to 26 September 2002. The DCM12,

installed on the seabed, recorded every 0.5 hours the current integrated in

5 cells each 11m thick, centred at 6.5, 13, 19.5, 26 and 32.5 meters. DCM12

and VMADCP data were rotated counter clockwise 30º to produce a

longitudinal and a transversal component referred to the main channel of

the ría.

Local winds were measured by the ship–mounted meteorological

station. Ship velocity was substracted to obtain the absolute wind velocity

with an error of ±2 m s–1. However, since local wind data have been

averaged over the three sector defined in the surface layer of the transect

(Figure 1), the standard deviation of the average wind velocity in each

sectors is ±2/N1/2 cm/s. Since N~7 the final SD is nearly 1 m/s.

Remote winds were taken from the Seawatch buoy of Puertos del

Estado off Cape Silleiro (Figure 1), a site representative for winds blowing

on the shelf outside the Ría de Vigo (Herrera et al. 2005). Local wind data

were analyzed to study the evolution of the transversal and the

longitudinal components of the wind. Low–pass remote winds are shown

for each period to define the prevailing oceanographic scenario in the ría

(Figure 2).

Total heat balances calculated following the equations of Laevastu

(1963) and Otto (1975) and freshwater fluxes (Ríos et al. 1992) were taken

from Piedracoba et al. (2005a, Chapter 2) for each sampling date.

Page 145: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

127

number Low/high fortnightly

date start end of

repet. tide tidal phase Semi-diurnal tidal

phase

14/02 ST 21/02 NT 18-2 11:06 16:12 27 12:47

21-2 10:54 17:06 28 15:40

21/02 NT

05/04 NT 13/04 ST 11-4 11:12 16:18 26 14:37

20/04 NT 22-4 8:54 15:09 31 11:42 27/04 ST

12/07 ST 18-7 9:11 14:58 25 15:01 19/07 NT

19/07 NT

22-7 8:59 14:44 27 12:54 25/07 ST

15/09 NT 19-9 9:53 15:12 28 15:02 22/09 ST

Table 1. Summary of relevant information for each sampling: date, starting and ending time of sampling, number of repetitions of the transect, time (GMT) of low or high tide, fortnightly tidal phase (spring tides-ST, neap tides-NT or transition between them) and scheme of semidiurnal tidal phase showing the change of tidal phase during every sampling.

-2-1012

24/0221/0218/02

-2-1012

24/0221/0218/02

-2-1012

20/0416/0412/04

-2-1012

20/0416/0412/04

-2-1012

27/0724/0721/0718/07

-2-1012

27/0724/0721/0718/07

-2-1012

25/0923/0921/0919/09

Page 146: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

128

-12-9-6-30369

12

15/2 18/2 21/2 24/202468101214

m3/sm/s

-12-9-6-30369

12

10/4 12/4 14/4 16/4 18/4 20/4 22/40

3

69

12

15

m3/sm/s

-12-9-6-30369

12

16/7 18/7 20/7 22/7 24/70

3

6

9

12

15m3/s

-12-9-6-30369

12

17/9 18/9 19/9 20/9 21/90

3

6

9

12

15m3/s

h

-30369

V (m

/s)

10:00 12:00 14:00 16:00-30369

10:00 12:00 14:00 16:00

V (m

/s)

19/09 19/09

-6-3036

09:00 11:00 13:00 15:00

V (m

/s)

-6-3036

V (m

/s)

09:00 11:00 13:00 15:00

-6-3036

V (m

/s)

09:00 11:00 13:00 15:00 -6-3036

V (m

/s)

09:00 11:00 13:00 15:00

18/07

18/07

22/07 22/07

d

-6-3036

V (m

/s)

12:00 14:00 16:00-6-3036

V (m

/s)

12:00 14:00 16:00

-6-3036

V (m

/s)

09:00 11:00 13:00 15:00-6-3036

V (m

/s)

09:00 11:00 13:00 15:00

11/04 11/04

22/04 22/04

a

g

Local wind longitudinal

E-W Shelf wind /QR

b

18/02

-12-9-6-30

12:00 14:00 16:00

V (m

/s)

-12-9-6-30

12:00 14:00 16:00

V (m

/s)

-12-9-6-30

11:00 13:00 15:00 17:00V

(m/s

)-12-9-6-30

11:00 13:00 15:00 17:00

V (m

/s)

18/02

21/02 21/02

transversal N-S

SS CS NS

c

e

Figure 2. Remote winds measured by the Seawatch buoy of Puertos del Estado for (a) February, (c) April, (e) July and (g) September 2002 (left axis) and continental runoff to the Ría de Vigo (right axis).Longitudinal and transversal local winds measured by the ship–mounted meteorological station during each sampling date: (b) 18/02 and 21/02; (d) 11/04 and 22/04; (f) 18/07 and 22/07; (h) 19/09. The average of the longitudinal (transversal) wind is represented for the same sectors where the longitudinal (transversal) residual currents were averaged. Black, grey and dotted lines correspond to the NS, CS and SS sectors showed in Figure 1. The triangle shows the time of low/high tide, when the tidal velocity is null.

Page 147: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

129

2.3. Limitations

The spatial coverage of the transect was limited in three ways.

Interference from the direct acoustic side–lobe return from the seabed

prevents reliable data being received from the bottom 15% of the water

column. Since the transducer is 2.5 m below the surface and the first bin

starts at 4 m from the transducer, the upper 6 meters are not sampled.

Further losses occur at the two ends of the section, where the ship turns

and perturbations of the ship gyrocompass may cause significant errors

(Pollard and Read 1989).

Current profiles were extrapolated to the bottom following the Von

Karman–Prandtl equation (Luek and Lou 1997):

0** zln

kuzln

ku)z(u −= (1)

where u(z) is the ensemble mean velocity (averaged to eliminate turbulent

fluctuations), z is the height above the seabed, k is Von Karman’s constant,

z0 is a length scale reflecting the bottom roughness, and u* is the friction

velocity that scales the turbulent velocity fluctuations. u* and z0 can be

estimated from the liner fit of u(z) vs ln(z). 2.4. Processing and accuracy

The quantitative use of ADCP data is limited by the accuracy of the

ship attitude measurements (positioning and heading) and the post–

acquisition data analysis. VMADCP measurements include the ship

velocity, SVr

. Therefore, SVr

has to be subtracted from the total velocity

measured by the VMADCP, TVr

, to obtain the current velocity CVr

:

δ+−=rrrr

STC VVV (2)

Page 148: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

130

where the parameter δr

includes i) the errors that affect TVr

(misalignment of the transducers, instrumental errors), and ii) the errors

that affect the auxiliary systems, Global Positioning System (GPS) and

gyrocompass.

The misalignment was corrected following the bottom track calibration

procedure described by Joyce (1989) and García Górriz et al. (1997). The

detected Doppler velocity has been rotated by a horizontal angle (α),

scaled up by 1+β due to a non–zero trim to the transducer/ship, and

added to the ship velocity to obtain the true water velocity. SVr

was

obtained from bottom tracking (BT) measurements. The error introduced

in SVr

by the GPS ( GPSδ ) is avoided since the velocity of the sea–bottom

relative to the ship are estimating from BT. These errors are the main

source of imprecision in the measurements (Pierce et al. 1988; García

Górriz et al. 1997) as SVr

is one order of magnitude larger than the current.

Ship heading info was provided by a Robertson RCG10 gyrocompass,

which used Synchro interface to allow the reception of headings

synchronized with acquisition. The gyrocompass error introduce a false

cross–track velocity component proportional to SVr

times the sine of the

error angle (Griffiths 1994). Additionally, there is an error inherent to all

gyrocompasses caused by the vessel course and, therefore, independent of

the type of gyrocompass. Summarizing, the instrumental and processing

error of an individual VMADCP measurement is together ± 4 cm/s.

However, when a large number (N) of individual VMADCP

measurements are averaged over a given sector of the transect, with

defined length and depth, the confidence limit of the average current

velocity, calculated as ±SD/√N, is much lower than the error of an

Page 149: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

131

individual VMADCP measurement because of the averaging effect of all

terms and properties involved in the calculations on the individual errors.

Since N>200 the final SD is at least one order of magnitude smaller.

2.5. Quality control

The VMADCP current velocity recorded at the mooring position was

compared with the DCM12 current velocity for all sampling dates.

DCM12 data of layers 2, 3 and 4 were considered. Layer 1 (average from 0

to 11 m depth) and layer 5 (average from 27 to 38 m depth) were not

compared because of the shadow zone of the VMADCP (reliable

measurements only between 6 m and 33 m depth). A linear regression

analysis applied to the longitudinal component of the current measured

by both instruments (Figure 3) indicated that the VMADCP overestimated

the velocity by 10% compared with the DCM12. Considering the errors in

obtaining the processed current data from both instruments, the

differences between them are insignificant and there was no detectable

difference between individual layer comparisons.

2.6. Residual versus total current

The first step to evaluate the residual currents from the VMADCP was

to extract the tidal current from each series of roughly 6 hours. It was

supposed that the time of slack waters coincide with the high or low tide,

therefore at this time the total current is equal to the residual. To test this

assumption, the DCM12 observed current at high/low tide was compared

with the DCM12 longitudinal component of the residual current averaged

over the sampling period. The residual current had been calculated by

filtering the observed DCM12 data with the A242·A25 filter (Godin 1972).

The differences were 1.6±0.1 cm/s in the surface and –1.8±0.2 cm/s in the

Page 150: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

132

bottom layers at high tide and –2.3±0.2 cm/s and –0.3±0.2 cm/s for

surface and bottom layers

VDCM12= 0.89(±0.02)VVMADCP

R2 = 0.83, n= 190, p< 0.0001

-10

0

10

20

30

-10 0 10 20 30VMADCP (cm/s)

1:1D

CM

12 (c

m/s

)

Figure 3. Comparison of the longitudinal current velocity between the ship–mounted ADCP and the moored DCM12.

at low tide. These differences were not considered because they were

lower than the error associated with the VMADCP individual

measurements. Since the tide is synchronous (Dyer 1997; Míguez 2003),

the VMADCP current at the time of high/low water can be considered as

the residual current along the transect. Thus for each sampling day the

time of high or low tide was sought in the more uniform DCM12 record

and the closest VMADCP survey was selected as the residual current in

the transect.

2.7. Numerical model

The model used in this work is a C++ version of the HAMSOM model,

developed by the Group of Physical Oceanography of the University of

Vigo (GOFUVI) from the core code of the Institut für Meereskunden

Page 151: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

133

(Hamburg). HAMSOM is a 3–D, z–coordinate (Arakawa–C grid),

baroclinic, semi-implicit finite–difference numerical prediction model. It

allows the introduction of salinity and temperature at any point of the

model, as well as the heat exchange with the atmosphere and the

evaporation/rainfall. The HAMSOM model has been recently adapted

and validated for the Ría de Vigo with field data at the outer (Souto et al.

2003) and inner zones (Diz et al. 2004) of the embayment.

For this study, an improved bathymetry was implemented from the

charts of the Instituto Hidrográfico de la Marina (Navy Hydrographic

Institute), with a higher horizontal resolution of 120x120 metres per model

cell. With this configuration, the transect studied divides in to 20 cells, a

number large enough to solve the lateral variability of the current velocity.

The domain of the model was set from 9º W to 8º 40’ W and 42º 5’ N to 42º

21’ N, wide enough to take into account the effect of shelf wind stress in

the ría. The model was forced with the wind data measured at the

Seawatch buoy. This wind data was applied uniformly outside of the ría,

while wind intensity and direction were changed inside the ría. On the

basis of historical observations in the ría, wind intensity was varied

linearly from its shelf value at the Cíes Islands to a fifth of this value at the

Rande Strait (Figure 1). If the angle between wind direction and the main

channel was less than π/3, then the wind direction was progressively

changed to align with the main channel direction of the ría. The river flow

was calculated from precipitation data as described by Rios et al. (1992b).

The model was initialized with all fields (current velocity, temperature

and salinity) homogeneously distributed all over the domain. Velocity

values were set to zero and salinity and temperature to the reference mean

values of the Ría de Vigo (S = 35.8 and T = 14 ºC). As the simulation

Page 152: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

134

evolves, the initial salinity and temperature relax towards the expected

values according to the fresh water discharge and the heat exchange with

the atmosphere. A relaxation time of 1 month was found to be long

enough to guarantee that the initial conditions had no influence in the

modeled values at the time of the first survey. The river flow was

introduced in the model through a Montecarlo method (see Souto et al.

2003), which adequately reproduces the fresh water flow relation to

salinity at the Rande Strait (Figure 1). A zero gradient condition was

applied at the 3 open boundaries of the model (Northern, Southern and

Western). An additional nudging condition was combined with the zero

gradient condition at the western boundary of the model, to properly

introduce the corresponding salinity and temperature of the upwelled

water. The nudging condition assumes that in an upwelling event,

temperature and salinity values at the bottom layer of the open boundary

will evolve towards the reference values of the Eastern North Atlantic

Central Water (ENACW). The salinity and temperature at the lower layers

of the western open boundary would be calculated as:

auFtFtFtF Rnn1n ⋅⋅−−=+ ))(()()( (3)

where F(tn+1) is the value of temperature and salinity for the next time

step, F(tn) is the actual field value, FR is the reference value corresponding

to the ENACW salinity and temperature at the corresponding depth, u the

velocity component perpendicular to the border and a is a constant whose

value was adjusted until the calculated salinity and temperature values

were similar to the observed values.

The turbulent closure scheme of Pohlmann (1996) was applied. It

proved simple enough to allow an acceptable timeframe in the

Page 153: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

135

simulations of the model and, at the same time, was accurate enough to

reproduce the vertical viscosity with an acceptable precision.

The data simulated by the model were vertically interpolated to obtain

the level of no movement. The transect was divided, as was done with the

field data, in five sectors, three of them above the level of no motion and

the other two below this level (see Results). Average currents were

calculated for each of these sections. The model was run once, starting on

20 January 2002, a month before the first survey, and ending with the last

survey. 3. Results

3.1. Morphology of the ría

The Ría de Vigo is a NE–SW orientated V–shaped coastal embayment

32 km long, which widens and deepens from the inner side (1 km wide, 20

m deep at the Strait of Rande) to the outer side (10 km wide, 50 km deep).

The ría is connected with the shelf through two mouths of different

morphology, separated by the Cies Islands: the northern mouth (2.5 km

wide and 25 m of maximum depth) and the southern one (5 km wide and

50 m of maximum depth). East of the Strait of Rande there is a shallow

bay (average depth, 3 m) known as San Simon Bay that collects most of

the continental water inputs to the ría.

Since the purpose of this work is to study the lateral variability of the

middle Ría de Vigo (V–shaped, 2.5 Km wide, 40 m maximum depth)

under different oceanographic conditions, five sectors ─three at the

surface (NS, CS and SS) and two at the bottom (NB and SB) ─ were

defined (Figure 1). The criterion to divide the surface and bottom sectors

was based on the level of no–horizontal motion for the current profiles,

the depth of the halocline for the salinity profiles and the depth of the

Page 154: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

136

thermocline for the temperature profiles, which varied depending on the

oceanographic conditions. The same sectors were defined for the

numerical model to compare modeled and observed average currents,

salinity and temperature. Local winds were also averaged in the same

three surface sectors.

In order to check if significant differences existed between the five

sectors a T–test was applied to the longitudinal component of the residual

velocity, salinity and temperature for each sampling period (Table 2).

3.2. Remote and local winds

The analysis of low–pass filtered shelf winds (Figures 2a, c, e, g) allows

a classification of the hydrodynamic situation of the ría according to their

intensity and direction. Contrary to the expected seasonal cycle, shelf

northerly winds were dominant during the winter surveys. On 18

February, the ría was at the peak of an intense upwelling episode

(northerly winds ranged from –8 to –10 m s–1) and, on 21 February, it was

under moderate upwelling (northerly winds from –7 to –8 m s–1). Two

upwelling episodes of different intensity were recorded in spring. On 11

April, the ría was at the peak of an intense upwelling (shelf northerly

winds of about –11 m/s) after a downwelling period. On 22 April, shelf

northerly winds reduced to –6 m/s, producing a moderate upwelling

episode. Moderate to weak shelf northerly winds were recorded during

the summer surveys. On 18 July, when shelf northerly winds ranged from

0 to –3 m/s, the relaxation of an intense upwelling episode that reached

the maximum intensity on 15 July was recorded. On 22 July, a spin down

situation was recorded. Finally, on 19 September, south easterly winds

blew on the shelf with velocities ranging from 3 to 5 m/s as part of a

downwelling episode that peaked on 18 September.

Page 155: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

137

The analysis of the longitudinal component of local wind (Figures 2b,

d, f, h) indicates that under upwelling conditions (moderate or intense) it

blew usually shelfwards, contributing to the water outflow promoted by

shelf northerly winds. This component was not homogenous along the

section; sometimes it decreased southward (on 18 and 21 February) and in

other cases it increased southward (on 11 April). On the other hand, the

longitudinal component of local winds blew inwards under upwelling

relaxation or downwelling conditions, favoring the inflow of shelf surface

water into the ría.

Regarding the transversal component of local winds (Figures 2b, d, f,

h), it is difficult to draw a consistent pattern. In general, the transversal

component was northerly, but when the local wind intensity was low,

transversal winds change from northerly in the northern sector to

southerly in the southern sector (on 18 and 22 July and 19 September

3.3. Residual currents

The analysis of the longitudinal component of the residual current in

the middle ría shows a positive 2–layered residual circulation pattern with

an outflow through the surface layer and an inflow through the bottom

layer under upwelling conditions (Figure 4a, c, e, g), while the circulation

pattern reversed to negative under downwelling conditions (Figure 4m).

Other circulation patterns recorded were a longitudinal circulation

scheme consisting of 2 layers in the central and southern sectors and a net

inflow through the northern sector during the summer upwelling

relaxation (Figure 4i) and a transient net entry of water characterized by

high velocities recorded during upwelling spin down (Figure 4k).

In general, there were significant differences in the longitudinal

component between the three surface sectors and also between the two

Page 156: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

138

longitudinal velocity transversal velocity

-5.7±2.1 -5.5±1.0 -13.4±2.0

5.5±1.0 4.5±0.6

-40

-30

-20

-10

0-3.7±0.9 -3.3±1.1 -1.5±0.8

2.8±0.8 3.5±0.7

-40

-30

-20

-10

0-6.3±2.7 -3.6±1.3 -2.7±1.8

-3.4±0.6-4.8±1.1

-40

-30

-20

-10

0 -6.8±2.5 -2.1±0.5

-4.8±0.9 -1.0±0.7

0.0±1.1

-40

-30

-20

-10

0

1.8±1.3

2.5±1.4 4.3±2.2

3.2±1.0

0.0±3.4

-4.9±1.4 -7.5±1.7

2.1±0.6 3.3±0.5

0.0±0.7

-3.8±3.0 -6.1±2.4 -8.9±3.6

5.7±1.0 3.5±1.6

-5.7±1.3 -8.2±0.9 -8.3±4.5

2.3±0.9 3.2±0.5

18/02/02

21/02/02

18/02/02

21/02/02

11/04/02 11/04/02

22/04/02 22/04/02 g h

a

e

b

c d

f

Figure 4. Longitudinal and transversal residual currents (cm s–1) for each sampling date. The section is viewed from outside the ría. The symbol represents an outflow (V < 0 or westwards) and represents an inflow (V > 0 or eastwards) of water into the ría. Solid lines show the LNM (level of no movement) along the transverse section. The arrows indicate the northward (V > 0) or southward (V < 0) lateral circulation.

Page 157: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

139

11.2±1.9 10.8±0.6 8.5±2.1

9.7±1.0 7.6±0.7

-40

-30

-20

-10

0-1.2±2.2

-6.6±1.0 -3.2±0.8

-2.3±1.2 2.0±0.6

-40

-30

-20

-10

0-5.9±1.4 -3.2±0.8

2.3±0.7

0.0±1.2

0.0±1.1

1000 1500 2000

-40

-30

-20

-10

0

2.0±1.1

-4.2±1.5 -3.8±1.1 -4.7±1.0

1.9±1.9

2.0±1.5 -3.3±1.7 -8.1±1.9

3.4±0.5 3.9±0.3

1000 1500 2000

-5.5±1.0 -5.1±0.8

1.1±1.1 2.1±0.9 4.7±1.4

18/07/02 18/07/02

22/07/02 22/07/02

19/09/02 19/09/02

i j

m

k l

n

Figure 4. (continued)

bottom sectors (Table 2). In fact, the surface outflow and inflow occurred

preferentially through the southern sector under upwelling (Figures 4a, e,

g, i) and downwelling (Figure 4m) conditions. Differences were also found

sometimes in the bottom sectors, resulting either in an asymmetric

distribution of maximum velocity between the surface outflow and

bottom inflow (Figure 4e) or maximum velocities of outflow/inflow

through the same side of the ría (Figure 4g). There are two cases (Figure

4i, 4m) in which significant lateral differences were found only in the

surface sectors, while the bottom layer presented a more homogenous

Page 158: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

140

distribution, and one case (Figure 4g) in which significant differences only

existed in the bottom sectors.

The transversal component of the surface residual current was

southward in all cases (Figure 4b, d, h, j, l, n) except one (Figure 4f), in

agreement with the dominant local northerly winds. A transverse

circulation cell with opposite surface and bottom flows developed in the

middle ría during moderate winter upwelling (Figure 4d), summer

upwelling relaxation (Figure 4j) and autumn downwelling (Figure 4n).

The lateral gradient in the intensity of the transversal components of local

wind and residual current velocity agreed with each other (Figure 4f, 4l).

The transversal component of the residual current velocity was of the

same order as the longitudinal component. The lateral differences in the

flow were not generally consistent with Earth’s rotational effects in the

North Hemisphere (the inflow shifted southwards and the outflow shifted

northwards).

Page 159: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

141

3.4. Salinity and temperature

The seasonal cycle of homogenisation/stratification has been described

by Nogueria et al. (1997) for the Ría de Vigo. During winter and early

spring, salinity controls the hydrographic variability of the Rías Baixas,

and a nearly homogenous distribution of temperature with thermal

inversion occurs.

In summer, salinity is more homogenous and thermal stratification

develops. Early in the autumn, haline and thermal stratification occurs

depending on the freshwater input and the heat exchange with the

atmosphere.

For the particular case of the study year, two upwelling events of

variable intensity promoted the entry of salty (>35.7) ENACW throughout

the bottom layer of the ría (Figures 5a, c) under the typical thermal

inversion conditions of February (Figures 5b, d). Significant spatial

differences were found in both thermohaline variables (Table 2).

Differences in the surface sectors were observed on 18 February as a

consequence of the fresher and colder river water and disappeared on 21

February as a consequence of the upwelling of ENACW to the surface

layer and the decrease of river flow (Figure 2a).

In April, marked haline stratification occurred because of continental

runoff, and the persistent upwelling of salty water (Figure 5e) caused

homogeneously high bottom salinities (Figure 5g). In addition, an abrupt

change was observed in temperature from a situation of thermal

homogenization on 11 April to the development of an incipient

thermocline on 22 April (Figure 5h), due to the increase of solar radiation,

585 cal·cm–2·d-1 (Piedracoba et al. 2005a, Chapter 2). The northern surface

sector was significantly saltier and warmer than the southern sector on 11

Page 160: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

142

April (Table 2), but no differences were found on 22 April. The bottom

layer presented homogenous distributions of salinity and temperature

both dates (Table 2).

velocity

date NS-CS CS-SS NS-SS NB-SB 18/02/2002 n.s * ** * 21/02/2002 * *** *** ** 11/04/2002 ** n.s *** *** 22/04/2002 * n.s n.s *** 18/07/2002 *** * *** n.s 22/07/2002 n.s * * *** 19/09/2002 *** *** *** n.s

salinity

date NS-CS CS-SS NS-SS NB-SB 18/02/2002 *** *** *** *** 21/02/2002 n.s n.s n.s n.s 11/04/2002 n.s n.s * n.s 22/04/2002 n.s ** * n.s 18/07/2002 ** *** *** *** 22/07/2002 n.s n.s n.s n.s 19/09/2002 n.s n.s n.s n.s

temperature

date NS-CS CS-SS NS-SS NB-SB 18/02/2002 *** ** *** *** 21/02/2002 n.s n.s n.s *** 11/04/2002 *** *** *** n.s 22/04/2002 n.s n.s n.s n.s 18/07/2002 n.s n.s n.s n.s 22/07/2002 n.s n.s n.s n.s 19/09/2002 *** n.s *** n.s

Table 2. t-TEST for (a) longitudinal residual current, (b) salinity and (c) temperature, between the three surface (NS, CS and SS) and the two bottom (NB and SB) sectors, for each sampling date; n.s: non-significant differences; * : p(0.1-0.05); ** : p(0.05-0.01); ***: p<0.01;

Page 161: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

143

In July, the halocline was difficult to discern (Figure 5i, k) and the

upwelling relaxation (on 18 July) and spin down (on 22 July) episodes

were characterized by high thermal stratification due to the combination

of the entry of cold water through the bottom layer and solar radiation in

the surface layer (Figures 5j, l). Significant differences in salinity were

found only during the upwelling relaxation episode of 18 July (Table 2).

In September, the thermohaline structure corresponded with a typical

autumn downwelling (Figures 5m, n). Shelf southerly winds produced a

deepening of isohalines and isothermes as a consequence of downwelling.

Salinity was homogenous, both vertically and horizontally, being difficult

to distinguish a halocline, but solar irradiation was able to maintain the

temperature gradient (irradiance of 171 cal·cm–2·d–1; Piedracoba et al.

2005a, Chapter 2), establishing a deep thermocline (from –15 to –17 m).

Significant temperature differences (p < 0.001) were found only between

the surface sectors (Table 2).

Page 162: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

144

salinity

-40

-30

-20

-10

013.28±0.03 13.31±0.06 13.32±0.06

13.67±0.04 13.70±0.02

N S35.14±0.03 35.30±0.08 35.44±0.05

35.71±0.04 35.78±0.02

-40

-30

-20

-10

0 N S13.21±0.04 13.38±0.07 13.50±0.04

13.66±0.03 13.74±0.02

35.46±0.03 35.48±0.05 35.48±0.04

35.78±0.03 35.78±0.01

34.82±0.05 34.76±0.09 34.68±0.04

35.35±0.04 35.24±0.08

-40

-30

-20

-10

013.68±0.00 13.56±0.02 13.51±0.01

13.58±0.02 13.58±0.02

35.19±0.05 35.17±0.04 35.25±0.02

35.56±0.04 35.59±0.02

-40

-30

-20

-10

014.66±0.14 14.62±0.07 14.76±0.11

13.21±0.17 13.17±0.15

18/02/02

21/02/02

11/04/02 11/04/02

22/04/02 22/04/02

18/02/02

21/02/02 c

a b

d

e

g

f

h

temperature ºC

Figure 5. Pairs of averaged salinity (left side) and temperature (ºC) (right side) for each sector every sampling date. Solid lines show the halocline (salinity) and thermocline (temperature), and dashed lines represent the LNM (also in Figure 4).

Page 163: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

145

35.42±0.01 35.46±0.03 35.50±0.02

35.55±0.00 35.57±0.00

-40

-30

-20

-10

014.77±0.29 14.51±0.40 14.45±0.32

12.33±0.13 12.35±0.15

35.43±0.01 35.42±0.01 35.43±0.00

35.51±0.02 35.49±0.02

-40

-30

-20

-10

016.36±0.32 16.22±0.40 16.50±0.32

13.56±0.46 13.39±0.62

1000 1500 2000

34.90±0.02 34.92±0.03 34.89±0.03

35.16±0.05 35.16±0.05

1000 1500 2000

-40

-30

-20

-10

017.01±0.05 16.87±0.03 16.88±0.03

16.49±0.15 16.43±0.15

18/07/02 18/07/02

22/07/02 22/07/02

19/09/02 19/09/02

k l

m n

i j

Figure 5. (continued)

3.5. Level of no movement, halocline and thermocline

The level of no movement (LNM), which separates the outflow and

inflow currents, was in most cases deeper than the halocline/thermocline

(Figure 5) with the exception of 21 February, when the

halocline/thermocline deepened more than LNM in the central sector

(Figures 4c, d). The LNM was located at an average depth of 10 m and

presented a variation of ± 4 m, depending on the upwelling/downwelling

intensity. In addition, the LNM was not homogenous along the section

(Figures 4, 5) and a variety of shapes were recorded in this study (positive

Page 164: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

146

and negative tilted, concave and convex). Different shapes were also

found for the halocline/thermocline (Figure 5). Two cases where no LNM

appeared were recorded. In one of them, on 18 July, there was no LNM at

the northern sector (Figures 4i, 4j). In the other, on 22 July (Figures 4k, 4l,

5k, 5l), the LNM was not established because of the transient net entry of

water along the whole transect (Figures 4k, l). Finally the deepest LNM,

halocline and thermocline were recorded during the September

downwelling, with a wide range of variation, from –8 to –24 m, and a

nearly analogous distribution of them all (Figures 4m, 5m, 5n).

4. Discussion

4.1. Factors controlling the lateral circulation in coastal embayments

and estuaries

The magnitudes of linear and non linear terms of the subtidal

momentum balance in the longitudinal and transversal directions was

analyzed for every survey to evaluate which are the main forcing

mechanisms responsible for the lateral variability of the Ría de Vigo

(Pritchard 1956; Dyer 1997). Assuming no time variations of velocity and

simplifying between non linear terms, the subtidal momentum balances

are (Valle Levinson et al. 2000; Cáceres et al. 2002; 2004):

)/()/()/( 22h

22z yuAzuAfvyuv ∂∂+∂∂=−∂∂ (4)

in the transversal direction and

)y/v(A

)z/v(A)y/p(gHfu)y/v(v22

h

22z

1

∂∂+

+∂∂+∂∂ρ−=+∂∂ −

(5)

in the longitudinal direction, where u is the longitudinal velocity, ν is

transversal velocity and, the brackets denote space and time tidal

averages.

Page 165: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

147

The two terms on the left of eqs. (4) and (5) represent the advective and

Coriolis accelerations, respectively, and the terms on the right represent

the pressure gradient accelerations (baroclinic and barotropic) and vertical

and horizontal frictions, respectively.

The coefficients Az and Ah denote the vertical and horizontal eddy

viscosities. For the vertical eddy viscosity wind stress was considered,

calculated from the average winds during the tidal cycle. The vertical

transfer of horizontal momentum from the atmosphere to the ocean is: 2

Da WCρ=τ (6)

where aρ = 1.3 kg/m3 is the density of air, W is the wind velocity and

DC is an empirical drag coefficient (dimensionless), 1.3 ·10–3 according to

Hidy (1972).

If zuAzs ∂∂ρ=τ / , then )//( zuA sz ∂∂ρτ= (Wong 1994; Csanady 1975).

Considering ρ= 1026.5 kg/m3 (the average density of the ría) and

0120zu ./ =∂∂ s–1 (average from the distributions of Figure 4), an estimate

of Az from wind stress will be 0.0068 m2/s. Finally, the relationship

proposed by Csanady (1975), f200uA 2z /*= , was applied to eliminate the

contribution of bottom friction, where u* is the frictional velocity (0.01

m/s) and f is the Coriolis parameter ( 151078.9 −−⋅ s ). This equation yields

a value of 0.0051 m2/s and subtracting this value from 0.0068 m2/s,

00170Az .= m2/s. For Ah, we assumed a value of 50 m2 s–1 on basis of the

two numerical models applied to the Ría de Vigo, HAMSOM and GHER

(GeoHydrodynamics and Environmental Research). HAMSOM has been

recently adapted and validated with observations at the outer (Souto et al.

2003) and inner (Diz et al. 2004) segments of the Ría de Vigo and in the Ría

de Ribadeo (Piedracoba et al. 2005b). GHER is a 3–D baroclinic model also

Page 166: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

148

adapted to the particularities of the Ría de Vigo by Torres–López et al.

(2001). In addition, the magnitude of Ah, is also consistent with the value

estimated for the Ría de Arousa by Otto (1975).

The value of Ah is arbitrary and it has never been measured in the

Galician Rías. For this reason, all the conclusions about the horizontal

friction terms are relative. Assuming an Ah, of 50 m2 s–1 the horizontal

friction terms result one order of magnitude higher than the Coriolis and

advective terms.

The relative magnitude of the momentum balance terms in the

transversal and longitudinal direction is presented in Figure 6 (note that

the momentum balance could not be calculated because all the baroclinic

and barotropic pressure gradients were not available). Magnitudes of the

terms in both directions were comparable (note that the magnitudes of u

and v were sometimes similar in Figure 4), highlighting the importance of

the lateral component under the funneling effect of the ría.

The Coriolis terms seem to be of the same order of magnitude than the

advective terms and one order of magnitude less than the friction or

baroclinic pressure gradient (calculated only in the transversal direction),

indicating that the residual circulation is dominated by non linear

frictional accelerations. These results are coherent with the schemes of

circulation presented in Figure 4, where the transverse circulation of some

scenarios seems to be opposite to the Earth’s rotation effects. For this

reason, it is reasonable to point to the wind as the main driving

mechanism of the mean flow during all the study cases and this forcing

means high non–linear effects, specially frictional influences. In general,

Earth’s rotation effects act in the ría but sometimes other non–linear

processes mask its action.

Page 167: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

149

18-2 21-2 11-4 22-4 18-7 22-7 19-90

5

10

15

20

25

30

m/s

2 (x10

-6)

Adv Cor Vfric Hfric

18-2 21-2 11-4 22-4 18-7 22-7 19-90

5

10

15

20

25

30

m/s

2 (x10

-6)

Adv Cor Vfric Hfric gPbar

(a) longitudinal

(b) transversal

Figure 6. Absolute values of the average longitudinal (a) and transversal (b) section momentum balance terms: advective, Coriolis, baroclinic pressure gradient (only calculated in b), vertical friction and horizontal friction terms.

Analyzing the seven study cases, the following classification arise

according to the lateral distribution of the longitudinal component of the

residual velocity as a function of shelf wind stress and topography:

Page 168: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

150

i) During the intense upwelling events of 18 February and 11

April, a lateral gradient of the longitudinal component of the surface

current was established in the middle ría. At those dates, the outflow

was intensified through the southern sector. However, whereas on

18 February insignificant differences were found between the two

bottom sectors, on 11 April the maximum residual bottom velocity

occurred through the northern sector, i.e. on the opposite side to the

strongest surface ouflow. The transversal component of the residual

velocity followed the local wind regime.

ii) During the moderate upwelling events of 21 February and 22

April, the lateral gradient of the longitudinal component of the

surface current was weaker compared with case (i). The lateral

gradient and magnitude of the longitudinal component of the

bottom residual current diminished too.

The net entry through the northern mouth of the ría under shelf

northerly winds could contribute to a stronger inflow through the

northern sector of the middle ría, as observed during the summer

upwelling relaxation of 18 July. On that date, the longitudinal

component of local wind, blowing at 2 to 6 m/s to the interior of the

ría, also contributed to the inflow.

iii) During the upwelling spin down event of 22 July, an intense

ingoing residual current was observed through all the sectors of the

middle ría as a response of the system to the sudden change of wind

forcing conditions. In this case, the largest velocities were recorded

through the northern sector.

Page 169: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

151

iv) During the downwelling event of 19 September, the reversed

residual circulation was stronger in the surface southern sector. This

lateral variability of the longitudinal component of the surface

residual current was because the net outflow through the northern

mouth of the ría, in response to the southerly winds that blew on the

shelf. This drainage by the northern mouth could weaken the surface

inflow through the northern sector of the middle ría.

Regarding the transversal component of the subtidal surface velocity, it

responded to the transversal component of local wind (direction and

gradient along the section), the effect weakening from surface to bottom.

Sometimes the ría did not respond immediately to the changes in local

wind, because the current pattern was dominated by shelf winds (see the

change of direction of the longitudinal local wind on 22 April). With the

exception of 11 April, the surface transversal velocity was always

southwards. A lateral gradient in the transversal component of the

subtidal circulation was found. This gradient was also found in the

evolution of local wind between sectors.

The typical 2–layered estuarine circulation pattern of the inner and

middle Ría de Vigo overlaps with the alongshore circulation of the outer

ría, which results from the interaction of shelf winds and the intricate

topography of the coast (Figure 1). The 3–D residual circulation pattern of

the outer ría inferred by Souto et al. (2003) seems to play a significant role

in the asymmetry found in the middle ría. Under upwelling conditions

(northerly winds on the shelf), an inflow of water occurs from surface to

bottom in the relatively shallow (<25 m) northern mouth of the ría Part of

this water feeds the surface outflow through the southern mouth and part

enters the ría, opposing the surface outflow associated with the upwelling

Page 170: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

152

circulation. The effect is stronger in the northern sector of the study

transect. The surface water outflow in the middle ría is retarded and so

lateral differences are generated in the longitudinal component of the

subtidal velocity. Moreover, the modulation of the currents by local winds

and their gradient across the channel are also responsible for differences

in the transversal component of the subtidal velocity. Under downwelling

conditions (southerly winds on the shelf), the coastal jet enters the ría

through its southern mouth and an outflow is observed from surface to

bottom in its northern mouth (Souto et al. 2003; Gilcoto 2004).

The momentum balance in partially stratified estuaries has been

assumed to be geostrophic in the transverse direction. This assumption

was made by Pritchard (1956) in the James River, where he did not

consider the influence of frictional terms, although they were comparable

with Coriolis accelerations. Fischer (1972) was among the first to suggest

that the most important residual mass transport mechanism in many

estuaries is the density–induced transverse circulation. Dyer (1997)

studied the relative importance of the terms of the lateral dynamical

balance in the Vellar estuary (salt wedge), Southampton water (partially

mixed) and Gironde estuary (well mixed). He concluded that there were

significant differences in the secondary circulation patterns and in the

relative importance of lateral terms in the dynamical balance and in the

salt transport between the three types of estuaries. In addition, he pointed

out the changes caused by the wind in the lateral circulation. He

confirmed that the most important terms in the lateral dynamic balance in

partially–mixed estuaries are the water slope, the internal density

structure and the centrifugal force, and Coriolis appears to be of

secondary importance, increasing its relevance in larger estuaries.

Page 171: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

153

Valle–Levinson and Atkinson (1999) found that non–linear advection

can be larger than Coriolis acceleration across the lower Chesapeake Bay,

thus invalidating the geostrophic assumption. Valle–Levinson et al. (2000)

found that the geostrophic approximation did not reflect the lateral

dynamics in shallow estuaries due to the importance of the friction and

advection terms. Cáceres et al. (2002) showed in a Chilean fjord how the

sign of the slope of the zero isotach (LNM) indicated that the lateral

momentum could not be explained by geostrophy.

In this study, we find that the advective terms are comparable with the

Coriolis term and that frictional terms are normally an order of magnitude

larger. In other words, non–linear effects contribute strongly to the

momentum balance and, are responsible for significant lateral circulation.

The study transect was located in a part of the ría where funneling and

bathymetric changes can generate increases of friction in the lateral

direction. The transverse baroclinic pressure gradient is generally one

order of magnitude larger than the Coriolis acceleration. Again, we

conclude that additional accelerations (friction and non linear advection)

are needed to produce the dynamic balance. In addition, even if the

vertical eddy viscosity is reduced by a half to eliminate the contribution of

bottom friction, the order of magnitude difference between the vertical

friction terms and the other terms prevails, indicating the importance of

the non–linear effects.

4.2. Numerical model outputs versus field observations

Field observations and numerical model outputs have been combined

to produce a view of the lateral circulation in the middle segment of the

coastal upwelling system of the Ría de Vigo. Modeled velocity, salinity

and temperature were compared with the VMADCP and CTD data for the

Page 172: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

154

seven study cases (Figure 7). The comparisons, averaged values over each

of the five sectors defined in Section 3, were used in every case. Root–

Mean–Square (RMS) differences between the observed and modeled

values and the corresponding determination coefficients (R2) are shown

for each variable and case.

Comparison between modeled and observed velocities yielded that

they are in agreement under upwelling conditions. In all the scenarios R2

was between 0.81 and 0.99 and RMS was lower than ~3 cm/s, i.e. less than

the processing error of VMADCP data. However, during summer

upwelling spin down (on 22 July), after a period of relaxation dominated

by wind calm the RMS velocity was particularly high and the

determination coefficient low (0.50). Thus, the net entry of water together

with the magnitude of the velocity were not reproduced by the model.

Concerning salinity, the model was able to simulate all the upwelling

cases. In general, R2 was between 0.7 and 0.99 and RMS was lower than

0.3, but no correlation was found in salinity during the spin down case (on

22 July). In addition, the RMS calculated for salinity was particularly high

on 19 September; the model tends to mix water in excess and the model

output shows an elevated salinity, homogenously distributed through the

water column.

With reference to the observed and modeled temperatures, the model

simulates adequately all the scenarios. R2 was between 0.91 and 0.99 and

RMS was lower than 1ºC except on 11 April, when no correlation was

found between simulated and observed temperatures, although the RMS

was within the instrumental error and, on 19 September, when R2 was 0.51

and the simulated was higher than the observed temperature.

Page 173: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

155

In most cases, the model reproduced the observations adequately. Only

when the ría was under upwelling relaxation or downwelling conditions

(22 July and 19 September) was the correlation and/or RMS between

observed and modeled data low and high, respectively.

The explanation to this was linked to the nudging condition imposed

only in the bottom layer of the western boundary of the model, and acting

only under upwelling conditions (Souto et al. 2003). In fact, the R2

between observed and modeled data considering the 7 cases were 0.59,

0.36 and 0.95 for velocity, salinity and temperature, respectively. But, if

the upwelling relaxation (22 July) and downwelling (19 September) cases

were eliminated the R2 were 0.83, 0.79 and 0.88 for velocity, salinity and

temperature, respectively.

In general, the model reproduces adequately the lateral variability of

the longitudinal residual current. Since the model was not run with true

local wind, the modeled transversal component of the residual current

was not presented. In any case, the field observations made in this study

indicate that the local wind acts as a secondary forcing acting on the

lateral variability of the longitudinal component of the residual velocity

and constitutes a primary forcing acting on the lateral variability of the

transversal component of the residual velocity.

Page 174: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

156

NS CS SS NB SB-20

0

2019/09

-20

0

2022/07

-20

0

2018/07

-20

0

2022/04

-20

0

2011/04

-20

0

20

21/02

-20

0

2018/02

34.5

35.0

35.5

36.0

12

15

18

34.5

35.0

35.5

36.0

12

15

18

34.5

35.0

35.5

36.0

12

15

18

34.5

35.0

35.5

36.0

12

15

18

34.5

35.0

35.5

36.0

34.5

35.0

35.5

36.0

12

15

18

NS CS SS NB SB34.5

35.0

35.5

36.0

NS CS SS NB SB12

15

18

12

15

18

R2 = 0.91 RMS = 2.39

R2 = 0.97 RMS = 0.13

R2 = 0.91 RMS = 0.06

R2 = 0.81 RMS = 2.54

R2 = 0.78 RMS = 0.10

R2 = 0.94 RMS = 0.15

R2 = 0.99 RMS = 4.37

R2 = 0.96 RMS = 3.16

R2 = 0.89 RMS = 2.25

R2 = 0.50 RMS = 8.50

R2 = 0.96 RMS = 2.50

R2 = 0.72 RMS = 0.34

R2 = 0.99 RMS = 0.11

R2 = 0.98 RMS = 0.12

R2 = 0.09 RMS = 0.26

R2 = 0.99 RMS = 0.59

R2 = 0.01 RMS = 0.33

R2 = 0.91 RMS = 0.42

R2 = 0.99 RMS = 0.40

R2 = 0.93 RMS = 0.88

R2 = 0.51 RMS = 0.46

salinity temperature Longitudinal

velocity

long

-V

Figure 7. Modeled and observed longitudinal velocity, salinity and temperature for the 5 sectors defined in the transverse section each sampling date. The root–mean–square (RMS) of the difference between observed and simulated data and the corresponding determination coefficients (R2) are shown.

Page 175: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

157

4.3. Importance of the lateral circulation in coastal embayments and

estuaries

The lateral variability of surface residual velocities has repercussions

on the energy associated with the mixing processes and create differences

in the salinity and temperature distribution along the transect. The

ENACW water that enters the ría preserved better its properties in the

southern than in the northern sector; in some cases, the northern sector

was fresher and warmer than the southern sector of the ría. This gradient

was caused by the Coriolis deflection of continental waters under a

positive residual circulation pattern and the same reason is applicable to

the preferential entry of the cold and salty ENACW through the southern

bottom sector (Piedracoba et al. 2005a, Chapter 2).

Previous works showed the lateral variability of salinity in the adjacent

Ría de Arousa. Lateral differences due to the Coriolis acceleration has

been observed before in this ría by Gómez Gallego (1971), Otto (1975) and

Castillejo and Lavín (1982), concluding that the northern shore is less

saline than the southern one, causing an asymmetry in the density

distribution (Otto 1975) and in the horizontal velocity field. Rosón et al.

(1997) highlighted the importance of introducing the Coriolis effect in the

Ría de Arousa by means of a non–stationary 2-D box model to determine

residual fluxes based on thermohaline properties. The influence of the

Coriolis acceleration in the salinity or the density fields has been studied

in a wide variety of systems. Some examples where lateral differences in

the density field were found are the James River estuary, a partially mixed

tributary to the Chesapeake Bay (Valle–Levinson et al. 2000), and some

fjords of southern Chile (Cáceres et al. 2004). Weiss et al. (1997) pointed

out that this lateral variability in physical and chemical properties is

Page 176: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

158

controlled by estuarine circulation. This was confirmed from lateral

variations in the pycnocline in Chesapeake Bay and the Neuse Estuary

(Reed et al. 2004) attributed to the buoyancy interactions from the upriver

nutrient–rich freshwater and the downriver high–salinity water. Reed et

al. (2004) pointed out that transverse and longitudinal differences in the

Neuse Estuary resulted from seasonal precipitation inputs, wind fields

and flow patterns, showing that not only rotational effects contribute to

the transverse differences. In practice, the flows involved in the

gravitational circulation ―fresher water driven by the river towards the

sea and an inflow of seawater near the bed together with the mixing of

these two types of water─ are not evenly distributed across the cross

section of estuaries.

Differences in the transversal component of the residual velocity forced

by wind stress did not contribute to the lateral homogenization of salinity

(temperature) and, at the same time, lateral differences in the density

structure also contribute to the secondary flows. In addition, since

upwelling and downwelling respond quickly to changes in wind speed

and direction, wind variations on timescales as short as a few hours are

able to generate lateral responses in estuaries (Reynolds–Fleming et al.

2004).

Piedracoba et al. (2005a, Chapter 2) demonstrated how remote shelf

winds controlled more that 65% of the temporal variability observed in

the subtidal circulation at the central station of the study section and that

local winds do not contribute significantly to this variability. However,

this work demonstrates that local wind effects are essential to resolve the

lateral variability of the transversal component of the residual circulation

of the middle Ría de Vigo.

Page 177: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Lateral variability in the Ría de Vigo

159

5. Conclusions

The lateral variability of the subtidal circulation of the coastal

upwelling system of the Ría de Vigo has been described under contrasting

hydrographic conditions, combining high resolution field observations

and a numerical model. The relative importance of the different sources of

lateral variability observed in the Ría de Vigo has been assessed. Three

main conclusions can be raised:

1) Significant lateral differences in the longitudinal and transversal

components of the subtidal velocity, salinity and temperature profiles

have been found in the central segment of the Ría de Vigo,

demonstrating the relevance of the secondary circulation to assess the

hydrodynamic characteristics of the ría under different oceanographic

conditions.

2) The secondary circulation of the Ría de Vigo can not be explained

attending only to rotational effects, because the deflection of currents is

opposite to the expected only by the Coriolis force in most of the

scenarios. In fact, the secondary circulation pattern depends on both

shelf and local winds. Shelf winds interact with the intricate

topography of the outer ría, consisting on a wide and deep southern

mouth and a narrow and shallow northern mouth separated by the

Cíes Islands. This interaction produces a conspicuous 3–D residual

circulation pattern in the outer ría that could be responsible for the

non–linear advection across the estuary, i.e. other accelerations,

advection and friction terms. The non–linear advection plays an

important role on the lateral dynamics of the ría, specially on the

longitudinal component of the subtidal velocity. On the other hand,

Page 178: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 3

160

local winds affected mainly the lateral variability of the transversal

component of the subtidal velocity.

3) The development of a circulation cell with opposite directions

between the surface and the bottom layers was generated by the

combined action of the transversal component of local winds and the

rotational effects.

Page 179: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4:

Origin and fate of a bloom of Skeletonema costatum during a winter upwelling/downwelling

sequence in the Ría de Vigo

Page 180: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 181: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

163

Chapter 4, the research work presented in this chapter is also a

contribution to the paper:

X. A. Álvarez-Salgado, M. Nieto-Cid, S. Piedracoba, B. G. Crespo, J.

Gago, S. Brea, I. G. Teixeira, F. G. Figueiras, G. Rosón, C. G. Castro, J. L.

Garrido, M. Gilcoto. 2005. Origin and fate of a bloom of Skeletonema

costatum during a winter upwelling/downwelling sequence in the Ría

de Vigo (NW Spain). J. Mar. Res. 63, 1127-1149.

Abstract: The onset, development and decay of a winter bloom of the

marine diatom Skeletonema costatum was monitored during a 10 d period

in the coastal upwelling system of the Ría de Vigo (NW Spain). The

succession of upwelling, relaxation and downwelling favourable coastal

winds with a frequency of 10–20 d is a common feature of the NW Iberian

shelf. The onset of the bloom occurred during an upwelling favourable ½

wk period under winter thermal inversion conditions. The subsequent ½

wk coastal wind relaxation period allowed development of the bloom

(gross primary production reached 8 g C m–2 d–1) utilizing nutrients

upwelled during the previous period. Finally, downwelling during the

following ½ wk period forced the decay of the bloom through a

combination of cell sinking and downward advection.

Page 182: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

164

Page 183: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

165

Resumen: Se siguió el inicio, desarrollo y finalización de un bloom de

invierno de la diatomea Skeletonema costatum durante un período de 10

días en el sistema de afloramiento de la Ría de Vigo (NO de España). La

sucesión de condiciones de afloramiento, relajación y hundimiento

siguiendo una frecuencia de 10-20 días de los vientos costeros es una

característica común de la plataforma del NO de la Península Ibérica. El

inicio del bloom tuvo lugar durante un periodo de 3-4 días de

afloramiento bajo condiciones de inversión térmica invernales. El

siguiente período de relajación de los vientos costeros de 3-4 días permitió

el desarrollo del bloom (la producción primaria bruta alcanzó valores de

8 g C m–2 d–1) a partir de los nutrientes aflorados durante el período

anterior. Finalmente, el período de hundimiento de 3-4 días forzó la caída

del bloom mediante la combinación de la sedimentación y la advección

vertical dirigida hacia la capa inferior.

Page 184: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

166

Page 185: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

167

1. Introduction

Understanding the origin, development, and fate of marine

phytoplankton blooms is a key scientific issue, either from the viewpoint

of the sustainable exploitation of marine living resources or the role

played by the ocean in the regulation of Earth climate. The topic becomes

specially relevant in ocean margins, which represent less than 8% of the

ocean surface area but comprise more than 25% of the ocean primary

production (Wollast 1998), up to 85% of the organic matter preserved in

marine sediments (Middelburg et al. 1993) and more than 90% of the fish

catches (FAO Fisheries Circular No. 920 FIRM/C920, 1997). Enhanced

nutrient fluxes from continental runoff, the atmosphere and the adjacent

open ocean (Walsh et al. 1991), together with an efficient nutrient

recycling based on a closed coupling between pelagic production and

benthic regeneration (Wollast 1993; 1998) are the reasons behind the high

productivity of the coastal zone. All these biogeochemical processes are

specially intensified in coastal upwelling regions because of the enhanced

entry of nutrient salts from the adjacent ocean (Walsh et al. 1991; Wollast

1998).

The typical annual succession of microplankton species in coastal

upwelling regions is characterised by diatom spring and dinoflagellate

autumn blooms delimiting a summer period with significant contribution

of heterotrophic components. The winter period is dominated by small

flagellate species. However, considerable temporal and spatial variability

affects this pattern in response to short–time–scale upwelling–relaxation

cycles (e.g. Blasco et al. 1980; Brink et al. 1980; Figueiras et al. 1994; Fermín

et al. 1996). Upwelling promotes diatom growth and relaxation favours

other species better adapted to stratified conditions, such as

Page 186: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

168

dinoflagellates (Margalef 1978). Consequently, upwelling causes sporadic

breaks of succession, returning to earlier stages. The combination of 1)

preceding phase of succession, 2) degree of water column stratification

and 3) wind intensity determine the stage to which succession is reset

(Estrada and Blasco 1979). In this sense, coastal upwelling must be

stronger during summer than in spring to cause a pioneer diatom bloom.

The annual cycle of chlorophyll concentration in NW Iberian shelf

waters is a mixture of two main components (Nogueira et al. 1997): 1) the

classical pattern of temperate ecosystems, characterised by spring diatom

and autumn dinoflagellate blooms separated by an unproductive winter

(light–limited) period and a summer (nutrient–limited) period dominated

by small flagellates in which heterotrophy is important (Figueiras and

Rios 1993); 2) short– time scale (10–20 d) successions from diatom to

flagellate populations in response to the conspicuous shelf wind–driven

upwelling/relaxation cycles that succeed during the upwelling favourable

season, from March to October (Pazos et al. 1995). It is remarkable that the

spring diatom bloom occurs at the time of the transition from the

downwelling to the upwelling favourable season, whereas the autumn

dinoflagellates bloom at the time of the transition from the upwelling to

the downwelling favourable season. The occurrence of upwelling or

downwelling conditions during the development of these blooms have

key implications for their origin, in situ growth versus accumulation, and

fate, off shelf export versus in situ respiration (Álvarez–Salgado et al.

2003). In any case, the annual cycle of shelf wind stress explains less than

15% of the total variability of the wind patterns off NW Spain, and more

than 70% of that variability concentrates at periods of less than one month,

being specially relevant the 10–20 d period of the upwelling/relaxation

Page 187: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

169

8.95º 8.90º 8.85º 8.80º 8.75º 8.70º 8.65º 8.60º

Longitude (W)

42.15º

42.20º

42.25º

42.30º

42.35ºLa

titud

e (N

)

Ría de Vigostn 00

Figure 1

Silleiro buoy

Cies Islands

riverOitabén

-50-20

-50-75

-100

episodes during the upwelling favourable season (Álvarez–Salgado et al.

2002).

Figure 1. Chart of the Ría de Vigo (NW Spain), indicating the position of the sampling site (stn 00), where the ADCP current meter was moored, the Seawatch buoy of Puertos del Estado off Cape Silleiro (in 350 m water) and the meteorological station of the airport of Vigo. The –20, –50, –75 and –100m isobaths are also included.

Although the short time scale hydrodynamic, chemical and biological

response to the meteorological forcing of the NW Iberian shelf is relatively

well characterised for the upwelling favourable season, a clear lack of

knowledge exists for the downwelling period. Short–time–scale wind–

driven upwelling events have been described during the downwelling

favourable season (Nogueira et al. 1997; Pardo et al. 2001; Álvarez et al.

2003), but studies about the impact on phytoplankton ecology and

biogeochemical cycles is restricted to their imprint in chlorophyll and

nutrient concentrations (Nogueira et al. 1997). The aim of this work is to

determine the origin and fate of a winter phytoplankton bloom of the

pioneer diatom S. costatum that occurred in the Ría de Vigo (NW Spain)

Page 188: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

170

during a 10 d period in February 2002. Meteorological, hydrodynamic and

hydrological (thermohaline, chemical and biological) field data and on

deck incubation experiments were combined to provide a realistic view of

the onset, development and abrupt termination of a winter bloom.

2. Methods

Sampling strategy. The time series station (stn 00), placed in the

middle segment of the Ría de Vigo (Figure 1) in 40 m water, was visited

about 1 hour before sunrise on 18, 21, 25 and 28 February 2002. Samples

were taken with a rosette sampler equipped with twelve 10–litre PVC

Niskin bottles with stainless–steel internal springs. Salinity and

temperature were recorded with a SBE 9/11 conductivity–temperature–

depth probe attached to the rosette sampler. Conductivity measurements

were converted into practical salinity scale values with the equation of

UNESCO (1985). Water samples for the analyses of dissolved oxygen,

nutrient salts, dissolved and suspended organic carbon, nitrogen,

phosphorus and carbohydrates, pigments and plankton counts were

collected from five depths: the surface (50% Photosynthesis Available

Radiation, PAR; 1.9±0.4 m), the depth of the 25% PAR (10±1 m), the depth

of the 1% PAR (18±1 m), 28±2m and the bottom (41±2 m).

Production/respiration rates, estimated by the oxygen incubation method,

were performed at the five levels too. Microzooplankton grazing

experiments were performed only at the surface (50% PAR).

Wind and current measurements. Shelf winds were measured by the

meteorological buoy of Puertos del Estado off Cabo Silleiro placed at

42.10N, 9.39W (Figure 1). An Aanderaa DCM12 Doppler Current Meter

was deployed on the seabed of the sampling site, at 40 m depth, for 24

days recording current velocity and direction at 30 minutes intervals. Five

Page 189: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

171

overlapped layers of 11 m depth were measured, centred at 6.5, 13.0, 19.5,

26.0 and 32.5 m approximately. Both subinertial winds and residual

currents (subtidal variability) were calculated by means of a A242A25 filter

(Godin 1972) with a cut off period of 30 hours, passed to the time series to

remove variability at tidal or higher frequencies.

Dissolved oxygen (O2). Samples were directly collected into calibrated

110 mL glass flasks and, after fixation with Cl2Mn and NaOH/NaI, they

were kept in the dark until analysis in the laboratory 24 h later. O2 was

determined by Winkler potentiometric end–point titration using a Titrino

720 analyser (Metrohm) with a precision of ±0.5 µmol kg–1.

Nutrient salts (NH4+, NO2–, NO3–, HPO42–, H4SiO4). Samples for

nutrient analysis were collected in 50–mL polyethylene bottles; they were

kept cold (4ºC) until analysis in the laboratory using standard segmented

flow analysis (SFA) procedures within 2 hours of their collection. The

precisions were ±0.02 µM for nitrite, ±0.1 µM for nitrate, ±0.05 µM for

ammonium, ±0.02 µM for phosphate and ±0.05 µM for silicate.

Dissolved organic carbon (DOC) and nitrogen (DON). Samples for

dissolved organic matter (DOM) were collected into 500 mL acid–cleaned

flasks and filtered through precombusted (450ºC, 4 h) 47 mm ø Whatman

GF/F filters in an acid–cleaned glass filtration system, under low N2 flow

pressure. Aliquots for the analysis of DOC/DON were collected into 10

mL precombusted (450ºC, 12 h) glass ampoules. After acidification with

H3PO4 to pH< 2, the ampoules were heat–sealed and stored in the dark at

4ºC until analysis. DOC and DON were measured simultaneously with a

nitrogen–specific Antek 7020 nitric oxide chemiluminescence detector

coupled in series with the carbon–specific Infra–red Gas Analyser of a

Shimadzu TOC–5000 organic carbon analyzer, as described in Álvarez–

Page 190: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

172

Salgado and Miller (1998). The system was standardized daily with a

mixture of potassium hydrogen phthalate and glycine. The concentrations

of DOC and total dissolved nitrogen (TDN) were determined by

subtracting the average peak area from the instrument blank area and

dividing by the slope of the standard curve. The precision of

measurements was ±0.7 µmol C L–1 for carbon and ±0.2 µmol N L–1 for

nitrogen. Their respective accuracies were tested daily with the

TOC/TDN reference materials provided by D. Hansell (Univ. of Miami).

We obtained an average concentration of 45.7 ± 1.6 µmol C L–1 and 21.3 ±

0.7 µmol N L–1 (n = 26) for the deep ocean reference (Sargasso Sea deep

water, 2600 m) minus blank reference materials. The nominal value for

TOC provided by the reference laboratory is 44.0 ± 1.5 µmol C L–1; a

consensus TDN value has not been supplied yet, but a mean±SD value of

22.1±0.8 µmol N L–1 for four HTCO systems and 21.4 µmol N L–1 for one

persulphate oxidation method has been provided by Sharp et al. (2004) as

a result of the Lewes intercalibration exercise. DON was obtained by

subtracting NT (=ammonium + nitrite + nitrate) to TDN.

Dissolved organic phosphorus (DOP). Samples were collected and

filtered as indicated for the DOC/DON samples. The filtrate was collected

into 50 mL polyethylene containers and frozen at –20ºC until analysis. It

was measured by means of the SFA system for phosphate, after oxidation

with Na2S2O8/borax and UV radiation (Armstrong et al. 1966). Only the

organic mono–phosphoric esters are analysed because poly–phosphates

are resistant to this oxidation procedure. Daily calibrations with

phosphate, phenyl phosphate and adenosine 5’–monophosphate (AMP) in

seawater were carried out. Standards of AMP were analysed in order to

Page 191: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

173

calculate the mono–phosphoric esters recovery (~80%). The precision of

the method was estimated as ±0.04 µmol P L–1.

Dissolved mono– and polysaccharides (MCho and PCho). Sampling

and storage procedures were identical than for DOP samples. MCHO and

PCHO were determined by the oxidation of the free reduced sugars with

2,4,6–tripyridyl–s–triazine (TPTZ) followed by spectrophotometric

detection at 595 nm (Myklestad et al. 1997). Quantification of MCHO and

total dissolved carbohydrates (d–CHO) was made by subtracting the

average peak height from the blank height, and dividing by the slope of

the standard curve (glucose). The estimated accuracy was ±0.6 µmol C L–1

for MCHO and ±0.7 µmol C L–1 for d–CHO, and the detection limit was ~2

µmol C L–1. See Nieto–Cid et al. (2004) for further details.

Particulate organic carbon and nitrogen (POC and PON). Suspended

organic matter was collected under low–vacuum on precombusted (450ºC,

4 h) 25–mm ø Whatman GF/F filters (POC/PON, 500 mL of seawater). All

filters were dried overnight and frozen (–20ºC) before analysis.

Measurements of POC and PON were carried out with a Perkin Elmer

2400 CHN analyzer. Filters were packed into 30 mm tin disks and injected

in the vertical quartz furnace of the analyser where combustion to CO2, N2

and H2O was performed at 900ºC. After separation of the gas products

into a chromatographic column, a conductivity detector quantified the C,

N and H content of the sample. Daily standards of acetanilide were

analysed. The precision of the method was ±0.3 µmol C L–1 and ±0.1 µmol

N L–1.

Particulate organic phosphorus (POP). The same procedures of

collection and storage than for POC/PON were followed, after filtration

of 250 mL seawater. It was determined by H2SO4/HClO4 digestion at

Page 192: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

174

220ºC of the particulate material collected over Whatman GF/F filters. The

phosphoric acid produced was analysed, after neutralisation, using the

SFA procedure for phosphate. Standards of phosphate were run every day

of analysis. The precision for the entire analysis was ±0.02 µmol P L–1.

Particulate carbohydrates (Cho). About 250 mL of seawater were

filtered and stored as indicated for POC, PON and POP. Cho

determination was carried out by the anthrone method (Ríos et al. 1998). It

is based in the quantitative reaction of sugars with anthrone in a strongly

acid medium at 90ºC, to give an intensely coloured compound. The

absorption was measured at 625 nm. The method was calibrated daily

with D–glucose standards. The estimated accuracy of the method was ±0.1

µmol C L–1.

Chlorophyll (Chl). One hundred mL of seawater was filtered through

GF/F filters and frozen (–20ºC) before analysis. Chl was determined with

a Turner Designs 10000R fluorometer after 90% acetone extraction

(Yentsch and Menzel 1963). The estimated precision was ±0.05 mg m–3.

Pigments. Sea water samples (1.5 L) were fractionated by filtering onto

a 47 mm diameter Millipore APFD filter (nominal pore size 2.7 µm) and

the filtrate filtered again through a Millipore APFF filter (nominal pore

size 0.7 µm). The filters were frozen at –80º C until analysis. Pigment

analyses were performed as described by Zapata et al. (2000): frozen filters

were extracted in 90% acetone and aliquots (140 µl) of clarified extracts

were mixed with of MilliQ water (60 µl) immediately before injection in a

Waters Alliance HPLC System (Milford, Massachusetts), comprising a

separations module, a photodiode array detector and a fluorescence

detector. The column (Symmetry C8 column, 150 x 4.6 mm, 3.5 µm particle

size) was themostatted at 25 ºC. Mobile phases were: A = methanol :

Page 193: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

175

acetonitrile : aqueous pyridine solution (0.25 M pyridine, pH adjusted to

5.0 with acetic acid) in the proportions 50:25:25 (v/v/v), and B =

acetonitrile : acetone (80:20 v/v). A segmented linear gradient was applied

(time, % B): 0 min, 0 %; 18 min, 40 %; 22 min, 100 %; 38 min, 100 %. Flow

rate was 1ml min–1. Pigments were detected by their absorbance at 440 nm

and in the case of chlorophylls by their fluorescence at 650 nm when

excited at 440 nm. Pigments were identified by co–chromatography with

authentic standards and by diode array spectroscopy. HPLC calibration

was performed using chlorophyll and carotenoid standards isolated from

microalgal cultures (see Zapata et al. 2000). The molar extinction

coefficients provided by Jeffrey (1997) were used for pigment

quantification.

Plankton counts. Plankton samples were preserved in Lugol’s iodine

and sedimented in composite sedimentation chambers. The sedimented

volume (10–50 ml) depended on the chlorophyll concentration of the

sample. The organisms were identified and counted to the species level,

when possible, using an inverted microscope. Small species were counted

along two transects at x250 and x400 magnification. The larger forms were

counted from the whole slide at x100 magnification. Counts are

representative for individuals > 5 µm.

Metabolic balance of the water column. Daily photosynthetic (Pg) and

respiration (R) rates of the microplankton community were estimated by

the oxygen light–dark bottle method (Strickland and Parsons 1972).

Samples collected before sunrise in 10 L Niskin bottles were transferred to

black polyethylene carboys. Five levels were sampled: 50%, 25% and 1%

of surface light, and two more depths below in the aphotic layer at 28±2m

and 41±2m. The carboys were gently shaken before sampling to prevent

Page 194: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

176

sedimentation of the particulate material. Series of eleven 110 mL Winkler

bottles composed of triplicate initial, and quadruplicate light and dark

subsamples were filled. Each series of eleven bottles were incubated for 24

hours (starting within 1 hour of the sun rise) in incubators placed in the

terrace of the base laboratory. The original temperature and light

conditions at each sampling depth were reproduced in the incubators by

mixing the appropriate proportions of the cold (10ºC) and warm (30ºC)

water provided by the cultivation facility of the base laboratory and using

increasing numbers of slides of a 1 mm mesh, respectively. Dissolved

oxygen was determined by Winkler potentiometric end–point titration as

indicated above.

Microzooplankton herbivory. Two experiments, on 21 and 28

February 2002, were carried out with surface water using the modified

dilution technique originally described by Landry and Hassett (1982). All

experimental containers, bottles, filters and tubing were soaked in 10%

HCl and rinsed with Milli Q water before each experiment. Water was

collected just before the CTD cast with a 30 L Niskin bottle dipped twice.

Water from the first dip was gravity filtered through a 0.2 µm Gelman

Suporcap filter and combined with unfiltered seawater from the second

dip in twelve 2.3 L clear polycarbonate bottles to obtain duplicates of 10,

20, 40, 60, 80 and 100% of plankton ambient levels. Bottles were

completely filled and incubated for 24 hours at the original light (50%

PAR) and temperature conditions in an incubator placed in the terrace of

the base laboratory. Initial and final replicate subsamples of 250 ml were

taken from all experimental bottles for the determination of chlorophyll

concentration. Chlorophyll samples were filtered onto 25 mm GF/F filters

and extracted with 90% acetone for 24 h at –20ºC. Chlorophyll

Page 195: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

177

concentration was determined by fluorometry using a Turner Designs

fluorometer. Changes in the chlorophyll concentration were used to

estimate the net growth rate (µ’= µ – g) of phytoplankton according to the

equation:

DgChlChl

lnt1

0

t ⋅−µ=⎟⎟⎠

⎞⎜⎜⎝

⎛⋅ (1)

where t is the duration of the experiment (24 h), Chl0 and Chlt are the

initial and final chlorophyll concentrations, respectively; µ and g are the

instantaneous rates of phytoplankton growth and grazing mortality,

respectively; and D is the dilution factor of the sample, which corresponds

to the relative concentration of the prey and predator population. µ and g

were estimated by linear regression of the daily net growth rate of

phytoplankton (µ’) against the dilution factor D.

Estimation of the chemical composition of biogenic materials. Fraga

et al. (1998) reviewed in great detail the average composition of plankton

carbohydrates (Cho), lipids (Lip), proteins (Prt), photosynthetic pigments

(Chl) and phosphorus compounds (Pho), which is summarised in Table 1.

This table also contains the relative contribution of each group to the

average composition of marine phytoplankton. From the average

composition of the biomolecules, the following set of five linear equations

can be written for the suspended organic material:

Chl46Pho45Lip53Cho17Prt139C ×+×+×+×+×= (2)

Chl52Pho76LipChoPrtH ×+×+×+×+×= 8928217 (3)

Chl5Pho31LipCho14PrtO ×+×+×+×+×= 645 (4)

Page 196: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

178

Chl4Pho12Prt39N ×+×+×= (5)

PhoP ×= 5 (6)

Chemical formula

% (w/w)

Phosphorus compounds C45H76O31N12P5 12

Proteins C139H217O45N39S 45.1

Chlorophylls a, b, c1 and c2

C46H52O5N4Mg 2

Lipids C53H89O6 16.5

Carbohydrates C6H10O5 24.4

Average composition C106H171O44N16P 100

Table 1. Chemical composition of the main organic products of synthesis and early degradation of marine phytoplankton according to Fraga et al. (1998). Percentages (in weight) of each group correspond to the average composition of marine phytoplankton.

Since suspended C, N, P, Cho and Chl have been measured, the system

can be solved to obtain the average chemical formula and the proportions

of the different biomolecules for each particular sample. Once the

biochemical composition of each sample is known, oxygen production

rates (Pg, in mmol O2 m–3 d–1) can be converted into chlorophyll

production rates (in mg Chl m–3 d–1) using the resultant RC= ∆O2/∆Corg

stoichiometric ratio and the Chl contribution to the suspended organic

carbon pool.

Page 197: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

179

3. Results

3.1. Water transports

Persistent north westerly winds of about –6 m s–1 blew on the shelf from

18 to 24 February and, then, they suddenly reversed to south easterlies

producing a peak of 6 m s–1 on 27 February 2002 (Figure 2a,b,c). This wind

regime produced a 2–layered positive residual circulation pattern with an

ingoing bottom current and an outgoing surface current from 18 to 24

February and a reversal of the flow from 24 to 28 February 2002 (Figure

2d). The circulation pattern had a strong impact on the salinity (Figure 2e)

and temperature (Figure 2f) distributions, characterised by the entry of

cold (<13.5 ºC) and salty (>35.8) Eastern North Atlantic Central Water

(ENACW) through the bottom layer from 18 to 24 February, which is

replaced by the colder (<13.2 ºC) and fresher (<35.4) shelf surface water

through the surface from 24 to 28 February 2002. The impact of shelf

winds on the thermohaline properties was substantially more decisive

than continental inputs (Figure 2b), quite reduced for this time of the year

(Nogueira et al. 1997), or heat exchange across the sea surface (Figure 2c).

3.2. Biochemical composition of the materials produced during the bloom

Ambient dissolved oxygen (Figure 2g), dissolved inorganic nitrogen

(Figure 2h) and silicate (Figure 2i) concentrations clearly indicated the

impact of the S. costatum bloom (Table 2) on the chemistry of the water

column with surface levels of 293 µmol O2 kg–1, 0.4 µmol N kg–1 and 0.94

µmol Si kg–1, respectively, during 25 February 2002 when typical ambient

concentrations for this time of the year are <250 µmol O2 kg–1, >4 µmol N

kg–1 and >5 µmol Si kg–1, respectively. More than 93% of the ambient

Page 198: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

180

Figure 2. Time evolution of shelf winds, in m s–1 (a), continental runoff, precipitation and evaporation, in m3 s–1 (b), heat balance, in cal cm–2 d–1 (c), residual currents, in cm s–1 (d), salinity (e), temperature, in ºC (f), dissolved oxygen, in µmol kg–1 (g), total inorganic nitrogen, in µmol k g–1 (h); silicate, in µmol k g–1 (i), Chlorophyll, in µg L–1 (j), fucoxanthin, in µg L–1 (k), suspended carbohydrates, in µmol L–1 (l), suspended organic carbon, in µmol L–1 (m), dissolved organic carbon , in µmol L–1 (n), dissolved carbohydrates, in µmol L–1 (o); gross primary production, in µmol kg–1 d–1 of O2 (p);→

-12-606

12

18/2 20/2 22/2 24/2 26/2 28/2

18 20 22 24 26 28

-10

0

10

20

30

40

QR- P

- E

Q P E

18 20 22 24 26 28 -250

-200

-150

-100

-50

0

H

-40

-30

-20

-10

0

g

j

m

p

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

dept

h (m

)de

pth

(m)

dept

h (m

)de

pth

(m)

18/2

21/2

25/2

28/2

d-40

-30

-20

-10

0

dept

h (m

)

-40

-30

-20

-10

0

h

k

n

q

-40

-30

-20

-10

0

-40

-30

-20

-10

0

-40

-30

-20

-10

0

18/2

21/2

25/2

28/2

e-40

-30

-20

-10

0a

bc

i

l

o

-40

-30

-20

-10

0

-40

-30

-20

-10

0

r

-40

-30

-20

-10

0

-40

-30

-20

-10

0

18/2

21/2

25/2

28/2

-40

-30

-20

-10

0

f

Page 199: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

181

→net primary production, µmol kg–1 d–1 of O2 (q); and respiration, µmol kg–1 d–1 of O2 (r). Panel b: black circles, conti nental runoff; grey circles, evaporation; open circles, precipitation.

dissolved inorganic nitrogen concentration during the study period was in

the form of nitrate.

Concomitantly, fluorometric chlorophyll levels (Figure 2j) increased from

>3 mg m–3 on 21 February to >14 mg m–3 on 25 February, four days later.

Surprisingly, on 28 February, three days later, surface levels reduced to 4

mg m–3 and maximum chlorophyll concentrations of >11 mg m–3 were

found in the bottom layer.

Diatoms and flagellates other than dinoflagellates were the main

components of the microplankton population, accounting for 62–97% and

3–36% of the total cell abundance respectively (Table 2). Dinoflagellates

and ciliates represented only a minor fraction. Diatoms were dominated

by the chain–forming species S. costatum that accounted for 85–97% of

diatom abundance, and small flagellates were almost the exclusive

component of other flagellates. Coinciding with the outgoing surface flow,

total cell numbers doubled, from a mean abundance of 275 cells mL–1 on

18 February to 565 cells mL–1 on 21 February.

This rate of increase, about twice for other flagellates (0.6 d–1) than for

diatoms (0.3 d–1), changed dramatically between 21 and 25 February when

total cell number increased by 30 times. During this period, when the

residual circulation stopped, diatoms accumulated at a rate of 11.7 d–1 (net

doubling rate of 1.4 d–1) considerably higher than for the other 3 groups

(0.36 ± 0.06 d–1). During the reversal of the residual circulation between 25

and 28 February, the microplankton population of the upper layer

decreased at a rate of –0.31 d–1, being –0.32 d–1 for diatoms and only –0.13

d–1 for other flagellates. The vertical distribution of diatoms in the photic

Page 200: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

182

layer was characterised by higher abundance at the surface than in the

bottom of the photic layer on days 18, 21 and 25, whereas they were

concentrated at 20 m (1% of incident light) on 28 February, where their

abundance was 6 times higher than in shallower depths.

18/02/2002 21/02/2002 25/02/2002 28/02/2002 Components

Diatoms 191 ± 188 (70%)

348 ± 215 (62%)

16598 ± 14807 (97%)

797 ± 851 (72%)

Dinoflagellates 10 ± 4 (4%) 12 ± 3 (2%) 32 ± 22 (0.2%) 31 ± 18 (3%)

Other flagellates

72 ± 23 (26%)

205 ± 106 (36%) 469 ± 294 (3%) 282 ± 121

(25%) Ciliates 1 ± 1 (0.3%) 1 ± 1 (0.2%) 2 ± 3 (0.01%) 3 ± 1 (0.3%)

Species/groups

Skeletonema costatum

176 ± 182 (91%)

321 ± 200 (92%)

16029 ± 14404 (97%)

674 ± 876 (85%)

Small Gymnodinium spp.

5 ± 2 (45%)

5 ± 4 (42%)

15 ± 25 (49%)

20 ± 13 (65%)

Small flagellates

68 ± 23 (94%)

193 ± 96 (94%)

432 ± 277 (92%)

266 ± 126 (94%)

Small oligotrichous ciliates

0.6 ± 0.6 (76%)

0.7 ± 0.7 (87%)

1.2 ± 2 (63%)

1.5 ± 0.1 (46%)

Table 2. Mean±SD cell abundance (cells mL–1) of the main microplankton components (diatoms, dinoflagellates, other flagellates and ciliates) and the dominant species or group of species within each component. In brackets are the contribution of each component to the total microplankton abundance and the contribution of each species/group of species to the component abundance.

The results on pigment composition during the sampling period

confirmed the predominance of diatom type pigments in the micro– and

nanoplankton fraction (operationally defined as organisms retained in the

Page 201: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

183

2.7 µm nominal pore size glass fibber filter, according to Rodríguez et al.

2003). Pigments in this fraction are characterized by high contents of

fucoxanthin (up to 3.77 µg L–1 in surface sample on 25 February) and

chlorophyll c2 (1.68 µg L–1 in the same sample) together with chlorophyll a.

Only surface samples taken on 25 February traced amounts of chlorophyll

c3 and 19’–hexanoyloxyfucoxanthin that indicate the presence of

flagellates belonging to the Division Haptophyta. Very small quantities of

chlorophyll b (ranging from 0.01 to 0.04 µg L–1) were detected on 25 and,

specially, on 28 February, suggesting that the decay of the diatom bloom

was followed by an increase in organisms representative of the Division

Chlorophyta (as reflected in the changes of proportions of diatoms and

flagellates that sampling day, see Table 2). In the fraction operationally

corresponding to picoplankton (organisms passing through 2.7 but

retained by 0.7 µm nominal pore size glass fibber filter; Rodríguez et al.

2003), only very small amounts of chlorophylls b and a were detected,

indicating low photosynthetic biomass. The occurrence of chlorophyll b

suggests the presence of cyanobacteria or, most probably and in

agreement with cell counts (“small flagellates” in Table 2), small sized

prasinophytes. This group was found to be abundant in this fraction in

late winter samples in the neighbour Ría de Pontevedra (Rodríguez et al.

2003).

The vertical distribution of pigments followed the development and

decay of the bloom, with chlorophyll c2, and specially fucoxanthin (Figure

2k), acting as good markers and paralleling the distribution of chlorophyll

a. This fact supports the consideration that, during the whole episode, the

main part of chlorophyll a was contributed by the same organism. During

the onset of the bloom, fucoxanthin amounts began to be important in

Page 202: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

184

surface samples on 21 February (0.45 and 0.42 µg L–1 for the two upper

samples, respectively). The development of the bloom (25 February) was

characterized by very high amounts of fucoxanthin near the surface,

which decreased dramatically in samples corresponding to 1 % PAR and

lower. The decay of the bloom reversed the situation, with highest values

of fucoxanthin measured at the lower depths sampled on 28 February. It

has to be noted that during the decay of the bloom, average concentrations

of each pigment in the water column did not suffer but slight variations

(fucoxanthin changed from 1.54 µg L–1 on 25 February to 1.45 µg L–1 on 28

February, chlorophyll c2 from 0.59 to 0.48 µg L–1 and chlorophyll a from

2.22 to 2.19 µg L–1). The smaller variation in the average concentration of

chlorophyll a is explained by the contribution of chlorophyll a from the

Chlorophytes responsible for the presence of chlorophyll b on 28

February. Pigment/chlorophyll a ratios measured in the upper layer

samples were the same as those measured in the lower layer samples

during the decay of the bloom. For the case of the

fucoxanthin/chlorophyll a ratio, it was 0.73, 0.68 and 0.65 at 50 % PAR, 25

% PAR and 1 % PAR on 25 February and 0.76, 0.68 and 0.65 at 1 % PAR,

28±2 m and 41±2 m, respectively, on 28 February.

The composition of the particulate organic material in the middle

segment of the ría, calculated with eqs 2 to 6, was characterised by a quasi

constant proportion of carbohydrates of about 11% (mol C/mol C) and

variable proportions of the most labile N– and P– rich materials and the

most refractory lipids: pigments, proteins and phosphorus compounds

proportions ranged from >60% (mol C/mol C) in the surface samples to

<40% (mol C/mol C) in the bottom samples (Figure 3).

Page 203: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

185

It contrasted with the composition of the particulate organic matter of

continental origin, with higher proportions of carbohydrates (about 20%

mol C/mol C). 1/η, the ratio of chlorophyll to dissolved oxygen

production in the upper layer, was calculated from the chemical formula

of particulate organic matter and the calculated amount of oxygen

produced during the synthesis of this material.

The ratio ranged from 3 to 9 mmol O2 mg Chl–1 (Table 3). The most

commonly used ∆O2/∆Corg ratio, also varied within a narrow interval

(1.42 to 1.43 mol O2 mol C–1), close to the Redfield value of 1.4 (Laws 1991;

Anderson 1995; Fraga 2001).

%lipids0 10 20 30 40 50 60

%N

/P

4

50

60

70

8020

30

40

50

60

70

0 20 40 60 80 1000

20

40

60

80

100

%CH

O

0

20

40

60

80

100

Figure 3. Ternary mixing diagram indicating the proportions of N and P compounds (chlorophyll + proteins + P compounds), carbohydrates and lipids in the particulate organic material of samples collected at stn 00. Solid dots, samples from stn 00; white dots, riverine samples.

Page 204: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

186

Date Chemical formula

18/02/2002 C95H154O30N12P 1.425 6.35 9.05

21/02/2002 C64H104O22N9P 1.421 3.96 5.63

25/02/2002 C82H132O30N13P 1.429 2.32 3.32

28/02/2002 C82H133O28N11P 1.426 2.87 4.09

Table 3. Average composition of the biogenic materials in the upper and lower layer of the middle segment of the Ría de Vigo during the sampling period in February 2002. The stoichiometric ratio of conversion of O2 changes into organic carbon changes, organic carbon changes into chlorophyll changes and O2 changes into chlorophyll changes are provided.

3.3. Production and consumption rates of Chla during the course of

the bloom

The time evolution of Pg rates (Figure 2p) is marked by the impressive

surface maximum of more than 110 mmol O2 m–3 d–1 on 25 February 2002,

during the transition from northerly to southerly shelf winds, which led to

an integrated primary production of 7.77 g C m –2 d–1 (Table 4). Since only

7.0% of the produced material was respired, 4.6% in the upper layer and

2.4% in the lower layer (Figure 2r), most of it (NCP = 7.26 g C m –2 d–1) was

available for transference to higher trophic levels (grazing) and/or

horizontal or vertical export (Figure 2q; Table 4). On the contrary, three

days later, after a brief downwelling episode (Figure 2a, d, e, f), Pg rates

had decreased to <20 mmol O2 m–3 d–1 and R rates kept relatively high,

specially at the bottom layer: 68.4 % of the produced material was

∆Corg∆OR 2

C =∆Chl∆Corg

∆Chl∆O2=

η1

C molO mol 2

Chl mgC mmol

Chl mgO mmol 2

Page 205: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

187

respired in the water column, 30% in the upper layer and 38.4% in the

lower layer (Table 4). This maximum of respiration at the bottom

coincided with high chlorophyll levels (Figure 2j).

In summary, 11.3% of Pg during the sampling period (18–28 February

2002) was respired in the upper layer, 12.1% was respired in the lower

layer and 76.6% was available for in situ grazing and horizontal

(advection) or vertical (advection, diffusion, sedimentation) export.

Table 4. Depth average gross (Pg) and net (NCP) primary production and respiration (R) rates in O2, carbon and chlorophyll units for the upper (u) and lower (l) layer of the middle segment of the Ría de Vigo during the sampling period in February 2002. The ∆O2/∆Corg and ∆Corg/∆Chl factors of Table 3 have been used to transform Pg rates from O2 to carbon and Chl units. A constant ∆O2/∆Corg= 1 mol O2 (mol C)–1 (equivalent to carbohydrate respiration) was used to transform R from O2 to carbon units.

Contrasting results were obtained from the two dilution experiments

conducted in the surface layer of the middle segment on 21 and 28

February to evaluate the microzooplankton grazing (Figure 4). On 21

February, i.e. during the onset of the S. costatum bloom, the growth rate

18-2

21-2

25-2

28-2

u l u l u l u l mmol O2 m-2d-1 71 0 138 0 925 0 201 0

g C m-2d-1 0.6 0 1.2 0 7.8 0 1.7 0 Pg mg Chl m-2d-1 8 0 25 0 279 0 49 0

mmol O2 m-2d-1 46 –34 115 –29 882 –22 141 –77

g C m-2d-1 0.3 –0.40 0.9 –0.35 7.3 –0.26 1 –0.93 NCP

mg Chl m-2d-1 8 0 25 0 279 0 49 0

mmol O2 m-2d-1 25 34 23 29 43 22 60 77 g C m-2d-1 0.3 0.4 0.3 0.35 0.5 0.3 0.7 0.93 R

mg Chl m-2d-1 0 0 0 0 0 0 0 0

Page 206: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

188

(average ± standard error) was 1.32±0.03 d–1 and the grazing rate 0.29±0.04

d–1.

On the contrary, on 28 February, i.e. during the abrupt decay of the bloom,

both the growth and grazing rates had reduced to a half of the values on

21 February: 0.64±0.01 and 0.12±0.01 d–1. Therefore, 20±3% of the daily

primary production of the surface layer (50% PAR) was grazed by

microheterotrophs. Assuming that this percentage is applicable to the

whole study period and that the species composition of microheterotrophs

and phytoplankton do not change with depth (Nogueira et al. 2000), a

constant with depth grazing rate was supposed. In addition, an average µ

for the photic layer <µ> was calculated from the empirical µ at the 50%

PAR and the proportion between the 50% PAR and the photic layer

integrated O2 primary production rate. The resultant <µ> was 0.69 d–1 on

21 February and 0.36 d–1 on 28 February. Then, the average Chl

concentration in the photic layer <Chl> as a result of <µ> and g was

calculated with the formula <Chl>= Chl0 (e(µ–g)t – 1)/(µ–g) (Calbet and

Landry, 2004). Finally, the photic layer integrated grazing rate, G = g

<Chl>, was estimated; It was 12.1 mg Chl m–2 d–1 (48% of Pg) on 21

February and 11.5 mg Chl m–2 d–1 (23% of Pg) on 28 February,

respectively. For the cases of 18 and 25 February, µ values were calculated

on basis of the initial Chl concentration and the gross O2 primary

production at the 50% PAR. Values of 0.57 and 1.13 d–1 were obtained,

respectively. The water column grazing rates (g), assumed to be 20% of

the surface µ, was 0.11 and 0.23 d–1 on 18 and 25 February, respectively.

Repeating the calculations performed for 21 and 28 February, it resulted

that <µ> = 0.33 d–1 and G = 2.44 mg Chl m–2 d–1 (20% of Pg) for 18

Page 207: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

189

February and <µ> = 0.63 d–1 and G = 51.4 mg Chl m–2 d–1 (23% of Pg) for

25 February.

dilution

0.0 0.2 0.4 0.6 0.8 1.0

ln(C

t/C

0)

0.00

0.25

0.50

0.75

1.00

1.25

1.50

dilution

0.0 0.2 0.4 0.6 0.8 1.0

Figure 4. Results of the grazing incubation experiments conducted on 21 (open circles) and 28 (solid circles) February 2002 with water of 10 m depth at stn 00. The logarithm of the ratio between the final and initial Chlorophyll concentration is represented against the dilution factor to obtain the corresponding phytoplankton growth (intercept) and microzooplankton grazing (slope) rates.

4. Discussion

A time series study of coastal winds and surface chlorophyll levels

revealed that winter upwelling events are relatively common in the rías of

the NW Iberian Peninsula and they used to be accompanied by

remarkable chlorophyll peaks (Nogueira et al. 1997). However, process

orientated studies of these winter upwelling events have not been

conducted until recently (Álvarez–Salgado et al. 2000; Pardo et al. 2001;

Álvarez et al. 2002) and they were just focused on hydrographic and

dynamic aspects. Only a biogeochemical study of the net ecosystem

production of dissolved organic carbon during a winter

upwelling/downwelling sequence has been reported up to now (Álvarez–

Page 208: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

190

Salgado et al. 2001). Therefore, this is the first comprehensive study of the

hydrodynamics and biogeochemistry of a winter phytoplankton bloom in

the NW Iberian upwelling system. To our knowledge, there are no

previous references to this phenomenon in other coastal upwelling

systems at comparable latitudes (off Oregon and off Chile), where

upwelling and downwelling favourable seasons can also be defined.

The average primary production of the Iberian shelf during the winter

time is <0.5 g m–2 d–1 from satellite and in situ estimates (Joint et al. 2002).

During the course of the bloom, a rate as high as 7.8 g C m–2 d–1 was

measured for an integrated Chl a concentration of 170 mg m–2 (surface

Chl a, 14 mg m–3) i.e. an assimilation number of 45 mg C (mg Chl)–1 . The

assimilation number dramatically decreased to 2.3 mg C (mg Chl)–1 during

the decay of the bloom (28 February). While an assimilation number of 2.3

mg C (mg Chl)–1 fall within the range of values obtained for the Ría de

Vigo (Tilstone et al. 1999) and for the adjacent shelf (Álvarez Salgado et al.

2003; Tilstone et al. 2003) during spring and summer, 45 mg C (mg Chl)–1

lies in the theoretical upper limit and, therefore, point to a high efficient

phytoplankton carbon fixation. The resultant primary production rate was

3 times the average primary production of the productive upwelling

season in the NW Iberian shelf: 2.5 g C m–2 d–1 (Arístegui et al. 2004) or

2.1–2.7 g C m–2 d–1 in the middle Ría de Vigo (Moncoiffé et al. 2000), a

number which can be considered representative for coastal upwelling

areas of the World Ocean (Wollast 1998).

Average surface Chl a levels at the same site during the month of

February from a time series of 9 years of measurements repeated twice a

week (n = 61 data) was 2.7 mg m–3, with 80% of the measurements within

the 0–6 mg m–3 interval (Nogueira et al. 1997). Therefore, surface Chl a

Page 209: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

191

levels on 25 February was five times the average and more than twice the

90% percentile. In fact, 14 mg m–3 represents the 90% percentile of Chl a

concentration during the productive upwelling season, from May to

October.

One of the possible fates of the produced material was respiration by

the community of micro organisms that occupy the middle segment of the

Ría de Vigo. In carbon units, microbial respiration was a relevant process

before the onset (18 February) and during the decay of the bloom (28

February), when 50% and 43% of the produced carbon was respired in the

photic layer, respectively. In fact, considering the respiration of the lower

layer too, it resulted that the middle segment of the ría was close to

balance (Pg = R) during those days. It is specially remarkable the high

respiration rate recorded in the bottom layer during the decay of the

bloom (0.9 g C m–2 d–1). For comparison, during the upwelling season,

Moncoiffé et al. (2002) obtained an average respiration of 0.5–0.6 g C m–2

d–1 in the aphotic layer of the middle segment of the Ría de Vigo. In any

case, considering the study 10 d period, it resulted that 77% of the

produced material was available for export to higher trophic levels, the

adjacent shelf or the sediments. This ratio of exportable versus produced

biogenic materials of 0.8, known as f– ratio after Eppley and Peterson

(1979), is characteristic of the most productive marine ecosystems. For

comparison, a mean f–ratio of 0.3 was obtained in the Ría de Vigo during

the upwelling season of 1991 (Moncoiffé et al. 2000) and 0.4 in the adjacent

Ría de Arousa during the upwelling season of 1989 (Álvarez–Salgado et

al. 1996). The f–ratio of a coastal upwelling system rarely exceeds 0.5

(Wollast 1998).

Page 210: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

192

A second possible fate is grazing by microzooplankton. Although

phytoplankton growth and grazing rates, estimated in this work from

dilution experiments, fell in the range of values reported for other

upwelling systems (Neuer and Cowles 1994; Landry et al. 1998; Edwards

et al. 1999), the relatively low grazing impact in the photic layer (21±3% of

phytoplankton growth, except on 18 February when it represented 48%)

contrasted with that found by Fileman and Burkill (2001) in shelf waters

of NW Iberian margin during summer, when microzooplankton

herbivory was an important cause of phytoplankton mortality. This could

be explained by the higher importance of microheterotrophs during

summer (Figueiras and Ríos 1993). In any case, microzooplankton

herbivory did not play a relevant role in the decay of the study winter S.

costatum bloom.

As a consequence of direct exudation and cell lysis during grazing

processes, part of the produced material is released to the photic layer in

the form of dissolved organic matter. During the development of the

bloom, from 21 to 25 February, the DOC concentration in the photic layer

increased from 65.0±0.5 to 70.8±0.5 mmol m–3, i.e. at 1.5±0.7 mmol m–3 d–1

or 0.2±0.1 g C m–2 d–1. The average DOC excess of the surface compared

with the bottom layer (∆DOC), 2.1±0.7 mmol m–3, times the average

surface outflow, 355±188 m3 s–1, yields a net offshore DOC flux of 9±4 g C

s–1. The surface outflow was calculated by multiplying the average surface

velocity (Figure 3c) times the corresponding cross section of the middle

segment of the ría. Assuming that this flux is representative for a free

surface area equivalent to the square of twice the internal Rossby radius of

deformation, 1100 m (Cusham–Roisin 1994), an extra DOC production of

0.16±0.07 g C m–2 d–1 would be necessary to balance the offshore flux.

Page 211: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

193

Therefore, neglecting other transport processes, the net DOC production

in the photic layer should be around 0.36±0.12 g C m–2 d–1 from February

21 to 25. This is <10% of the average Pg, as expected under mesotrophic

conditions (Teira et al. 2001a; 2001b). In addition, <30% of ∆DOC was

carbohydrates, a pattern that is also characteristic of nutrient–replete

conditions (Normann et al. 1995). By contrast, during the decay of the

bloom, the DOC concentration in the photic layer decreased from 70.8±0.5

to 67.5±0.5 mmol m–3, i.e. at –1.1±0.7 mmol m–3 d–1 or –0.2±0.1 g C m–2 d–1.

This situation is typical of nutrient–deplete conditions, with

microheterotrophs taking advantage of the situation. Net consumption of

dissolved organic matter under downwelling conditions has been

previously described in the Iberian upwelling system by Álvarez–Salgado

et al. (2001). The absence of nutrients produced that >70% of the DOC

decrease in the surface layer were dissolved carbohydrates, a situation

also typical of oligotrophic conditions (Normann et al. 1995).

A forth possible fate is the downward transport to the aphotic layer

and the sediments. Considering the average Chl concentration of the

photic layer (7.20±0.05 mg m–3) and the downward convective velocity

(10±2 m d–1), obtained from the time evolution of the depth of the 35.5

isohaline, between 25 and 28 February, it results in a downward transport

of 72±12 mg Chl m–2 d–1. This is 26±2% of the average phytoplankton

production (164 mg Chl m–2 d–1) plus Chl import from the outer ría (88±15

mg Chl m–2 d–1) plus Chl loss in the photic layer (23.6±0.4 mg Chl m–2 d–1)

from 25 to 28 February. The Chl import was obtained multiplying the

average Chl concentration of the photic layer times the average surface

inflow, 688±255 m3 s–1, and dividing by a surface area equivalent to the

square of twice the internal Rossby radius of deformation. The Chl loss

Page 212: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

194

was calculated by subtracting the average concentration on 28 February

from the average concentration on 25 February. Microzooplankton

grazing represented 35±5 mg Chl m–2 d–1, i.e. 13±2% of the total Chl

change in the middle ría. Assuming again that the downward convective

flux is representative for a surface area equivalent to the square of twice

the internal Rossby radius of deformation, it results that the surface

inward flow was not significantly different from the vertical downward

flow. Therefore, sedimentation would be 169±20 mg Chl m–2 d–1; this is

61±5% of the downward transport of Chl, at an average sedimentation

rate of 15±3 m d–1. This rate is within the range of variability of

phytoplankton living and dead cells found in the literature (Druon and Le

Fèvre 1999). In a mesocosms study simulating a diatom winter/spring

bloom, Riebesell (1989) obtained that the sedimentation rate of S. costatum

increased from <0.5 m d–1 during the onset to >4 m d–1 during the decay.

By contrast, other diatoms present in the mesocosm of Riebesell (1989),

such as Thalassiosira spp. and Leptocylindrus minimus sank at rates always

<2 m d–1. In any case, maximum sinking rates of S. costatum in the

laboratory are 1.4 m d–1 (Smayda and Boleyn 1966), which suggests that

the high sinking rates measured in the field resulted from the formation of

aggregates that increases in abundance during the decay of blooms

(Riebesell 1989), under condition of high phytoplankton biomass and

temperature increase (Thronton and Thake 1998).

The success of a phytoplankton population depends on its ability to

strike a favourable balance between rates of cell division, grazing losses

and sinking (Pitcher et al. 1989). Comparison of the relevance of physical

versus biogeochemical processes for the rapid decay of the S. costatum

bloom revealed that sinking was the main mechanism removing cells

Page 213: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

195

from the photic layer in the ría. However, this high sedimentation rate

was accompanied by an abrupt transition from upwelling to downwelling

conditions; whereas upwelling promotes nutrient salts from the lower to

the upper layer, this key fertilization mechanism is depressed under

downwelling conditions. Therefore, nutrients depleted (NO3– < 0.4 µM N)

during the blockage of the residual circulation in the middle ría from 21 to

25 February, were not replenished from 25 to 28 February. The

physiological mechanisms responsible for the regulation of phytoplankton

buoyancy seem to be determined by ambient light intensity and nutrient

regime (Bienfang 1981; Johnson and Smith 1986; Culver and Smith 1989).

Therefore, the absence of nutrients was the probable cause of the large

sedimentation rates from 25 to 28 February as observed in other diatom

blooms (Smayda 1970; Richardson and Cullen 1995; Waite et al. 1992).

The later authors found that threshold nitrate concentrations

approximating KS values signalled the initiation of increased

sedimentation. For the case of S. costatum, the threshold ambient nitrate

level was 1 µM. Consequently, there seems to be a physically mediated

biogeochemical control of the decay of the study winter S. costatum bloom.

Winter blooms are very relevant for the metabolic balance of the

ecosystem because they represent a huge entry of freshly produced

biogenic materials to the pelagic and benthic communities of the ría

during a time when the primary production is commonly at low rates. In

the study case, 70% of the material produced during the 10 d period was

transferred to the lower layer of the ría. Only 20% of this material was

respired by the heterotrophic microbial community of the aphotic layer

and the remaining 80% would eventually feed the benthic communities. It

is important to note that the fate of the bloom would be completely

Page 214: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

196

different if the period of calm would be followed by a new upwelling

event. In that particular case, most of the material produced in the middle

ría would be exported to the adjacent shelf and, eventually, would feed

the benthic communities there. Furthermore, no pigment degradation

products, apart from small amounts of chlorophyllide a, were observed

during the decay of the bloom. Pigment/chlorophyll a ratios were the

same in the photic layer during the development of the bloom as in cells

sunk to the lower layer during the decay, strongly supporting that

chloroplasts still keep their organization at this stage. It suggests that S.

costatum cells could be still viable and, thus, able to bloom again if a

immediate strong upwelling event (able to promote the sunk cell to the

photic layer) succeeded after 28 February.

The biochemical composition of the produced materials is an excellent

indicator of its nutritional quality for the pelagic and benthic

communities. N and P compounds are preferentially used by

microheterotrophs as substrates for structural and reproductive purposes.

The average proportion of N and P compounds produced in the upper

layer was 52% during the 10 d study period, this is about the average

composition of marine phytoplankton (Laws 1991; Anderson 1995; Fraga

2001). This proportion reduced to 47% in the aphotic layer, where the

most refractory photosynthates, lipids, represented as much as 46%. This

is the result of the preferential consumption of N and P compounds by the

water column heterotrophs of the aphotic layer. Moreover, the material

that arrived to the pelagic sediments is still of high quality, with proteins

and phosphorus compounds representing 44% of the carbon deposited.

Page 215: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Skeletonema costatum winter bloom

197

5. Conclusions

The extremely variable hydrographic conditions of the western coast of

the Iberian Peninsula, characterized by a succession of shelf wind

stress/relaxation events, allow the unusual development of massive

phytoplankton blooms during the winter time. In 10 d, the onset,

development and decay of pioneer diatoms can occur, if the appropriate

combination of upwelling (onset), relaxation (development) and

downwelling (decay) conditions succeed. The downward transport of

these winter blooms constitute the main entry of fresh materials to the

benthic communities during the winter time.

Page 216: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 4

198

Page 217: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5:

Physical-biological coupling in the

coastal upwelling system of the Ría de Vigo :

An in situ approach

Page 218: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 219: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

201

Chapter 5, the research work presented in this chapter is also a

contribution to the paper:

S. Piedracoba, M. Nieto-Cid, C. Souto, M. Gilcoto, G. Rosón,

X. A. Álvarez-Salgado, R. Varela, F. G. Figueiras. 2005. Physical-

biological coupling in the coastal upwelling system of the Ría de Vigo

(NW Spain). I: An in situ approach. Mar. Ecol. Prog. Ser. (submitted).

Abstract: The fate of the inorganic and organic N trapped in the coastal

upwelling system of Ría de Vigo (NW Spain), accumulation/export versus

production/ consumption, was studied at the short time–scale (2–4 d)

during July and September 2002. A transient geochemical box model was

applied to the measured residual currents and concentrations of inorganic

(NT), dissolved (DON) and particulate (PON) organic N to obtain the i)

net balance of inputs minus outputs (i – o); ii) the net accumulation

(V·dN/dt); and iii) the net ecosystem production (NEP) of NT, DON and

PON. Previously unaccounted lateral variations in residual currents and

N species concentrations in the middle ría were considered. The average

NEP during July (107 mg N m-2 d-1) indicates an autotrophic metabolism

of the ría. About 25% of this material was exported to the shelf and the

remaining 75% was transferred to the sediments or promoted to higher

trophic levels. During strong and weak summer upwelling episodes, from

30 to 70% of the organic N exported to the shelf came from materials

previously accumulated in the ría. During summer relaxation, 60-70% of

the accumulated DON and PON came from in situ consumption of NT and

the remaining 30-40% was allocthonous. In September, the metabolism

changed from heterotrophic to slightly autotrophic (NEP = 26 mg N m-2

d-1). Preferential consumption of imported DON occurred during autumn

Page 220: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

202

downwelling conditions whereas imported PON accumulated in the ría.

During autumn relaxation, 100% of the DON and 50% of the PON

accumulated in the study volume was produced in situ at the expenses of

N nutrients previously accumulated into the system. During weak

upwelling conditions, DON accumulated in the ría: 42% imported from

the shelf and 58% produced in situ. In contrast, PON losses were due to

consumption (42%) and export (58%).

Page 221: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

203

Resumen: Se estudió el destino del nitrógeno inorgánico y orgánico en

el sistema de afloramiento de la Ría de Vigo (NO de España) a corta escala

temporal (2-4 días) durante los meses de julio y septiembre de 2002, con

el fin de valorar la importancia de la acumulación y la exportación frente a

los procesos de producción y consumo. Para ello se aplicó un balance

geoquímico a partir de medidas de corriente residual y concentración de

nitrógeno inorgánico disuelto (NT), nitrógeno orgánico disuelto (DON) y

nitrógeno orgánico particulado (PON), para obtener: i) el balance neto de

entradas menos salidas (i – o); ii) la acumulación neta (V·dN/dt); y la

producción neta del ecosistema (NEP) de NT, DON y PON. Se han tenido

en cuenta las variaciones laterales en la corriente residual y en las

concentraciones de los compuestos de nitrógeno. La NEP promedio del

mes de julio (107 mg N m-2 d-1) indica un metabolismo autotrófico. Sobre

el 25% se exportó a la plataforma y el 75% restante sedimentó o se

transfirió a niveles tróficos superiores. Entre el 30% y el 70% del nitrógeno

orgánico que se exportó a la plataforma procedía de material previamente

acumulado en el sistema durante los episodios de aforamiento intenso y

moderado de julio. A lo largo del período de relajación de julio, el 60-70%

del DON y el PON acumulado en el sistema venía del consumo in situ del

NT y el 30-40% restante procedía del exterior del sistema. En septiembre,

el estado metabólico cambió de heterotrófico a ligeramente autotrófico

(NEP = 26 mg N m-2 d-1). Durante el período de hundimiento de

septiembre, se consumió preferentemente el DON importado del exterior,

mientras que el PON importado se acumulaba en la ría. Sin embargo, bajo

condiciones de afloramiento moderado de septiembre, el DON se

acumuló en la ría: el 42% procedía de la importación desde la plataforma

Page 222: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

204

y el 58% de la producción in situ. En cambio, en el sistema había pérdidas

netas de PON, tanto por consumo (42%) como por exportación (58%).

Page 223: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

205

1. Introduction

The concept of new production (NP), i.e. the fraction of the primary

production (Pg) that is supported by allocthonous nutrient salts (Dugdale

and Goering 1967), is a key variable to understand marine ecosystems

functioning either form the viewpoint of the exploitation of marine

resources or the global climate change. On one side, NP represents the

threshold for biomass exploitation without affecting the integrity of a

marine ecosystem at the long term (Quiñones and Platt 1991). On the

other side, NP quantifies the efficiency of the biological pump as a trap for

the antropogenic CO2 accumulated in the atmosphere (Longhurst 1991).

The most straightforward way of calculating NP is considering the

balance of nutrient input minus outputs in the ecosystem under

consideration (Smith and Hollibaugh 1993; Wollast 1993; 1998). However,

this conceptually simple method involves operational difficulties: velocity

and turbulent diffusion profiles, together with nutrient concentrations,

have to be determined with a frequency higher than the flushing time of

the study system. For this reason, different indirect methods have been

developed over the last decades to evaluate NP.

Gravity sinking and subsequent mineralization is the main fate of NP

in vast areas of the open ocean, where the horizontal circulation is quite

slow and most of Pg is processed within the microbial loop (Legendre and

Rassoulzadegan 1995; Cotner and Biddanda 2002). Therefore, assuming

steady state conditions, NP should be equal to the sedimentation rate

measured at the base of the photic layer (Eppley and Peterson 1979). In

vitro approaches has also been used to estimate NP at these open ocean

areas such as the 15N labelled nitrate/ammonium+urea method developed

by Dugdale and Goering (1967) or the classical light–dark oxygen

Page 224: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

206

incubation method (Strickland and Parsons 1972), assuming that the net

community production (NCP) estimated with the light bottle can be

equalled to the NP (Quiñones and Platt 1991). In fact, recent application of

the late method in subtropical gyres has led to a great controversy on the

autotrophic (NCP > 0) or heterotrophic (NCP < 0) status of these systems

(del Giorgio et al. 1997, Duarte and Agustí 1998, Williams 1998, Serret et

al. 2001; del Girogio and Duarte 2002). In any case, the validity of shallow

sediment traps (Buesseler 1991; Ducklow et al. 1995), 15N labelled

incubations (Legendre and Gosselin 1989, Wafar et al. 1995) and light–

dark oxygen incubations (Hansell et al. 2004) to estimate NP has been

openly challenged.

The evaluation of NP is more complex in coastal systems, where

horizontal advection is very relevant and a significant fraction of Pg is

transferred to the metazoans (Legendre and Rassoulzadegan 1995; Cotner

and Biddanda 2002). In the particular case of coastal upwelling systems at

temperate latitudes, the large flushing times and the non steady state

conditions (Alvarez Salgado et al. 1996; Smith and Hollibaugh 1997), adds

more complexity to the estimation of NP with a mass balance approach

and either sediment traps and/or in vitro incubations could became poor

approximations to the NP of these ecosystems (Smith and Hollibaugh

1997; Alvarez-Salgado et al. 2001).

In this work and its companion paper (Piedracoba et al this issue), we

deal with the estimation of NP in the coastal upwelling system of the Ría

de Vigo (NW Spain) under contrasting upwelling (in July) and

downwelling (in September) conditions comparing and/or combining the

mass balance, the sediment trap and the in vitro incubation approaches.

Part I is devoted to the estimation of the balance of input minus outputs of

Page 225: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

207

nitrogen species at the short time scale (< 4 days) by combining dynamic

and chemical determinations along a transverse section in the central

segment of the Ría de Vigo. A mass balance box model has been

developed to evaluate the relative importance of net accumulation and net

ecosystem production (NEP = NP) as mechanism of organic and inorganic

nitrogen trapping in this ecosystem.

2. Methods

Survey area. The Ría de Vigo is a large (2.76 km3) and V-shaped

indentation in the western coast of the Iberian Peninsula (Fig. 1). It

experiences wind-driven upwelling of cold, salty, nutrient and CO2 rich

Eastern North Atlantic Central Water (ENACW) from April-May to

September-October (Wooster et al. 1976; Fraga 1981). Upwelling events

occur with a 1-2 wk frequency (Blanton et al. 1987, Álvarez-Salgado et al.

1993). Downwelling is the dominant process the rest of the year; it is

characterised by the advection of warm, nutrient and CO2 poor shelf

surface water to the ría (Castro et al. 1997). The main tributary to the ría is

the river Oitabén-Verdugo, which drains an average annual flow of 15 m3

s-1 (Nogueira et al. 1997) into San Simón bay, a semi enclosed basin of ~ 20

km2 (Figure 1).

Sampling strategy. Five stations in the middle segment of the Ría de

Vigo (41±2 m max depth) were sampled about 1 hour before sunrise on 15,

18, 22 and 26 July and 17, 19, 23 and 26 September 2002. Water samples

from stn 00 (Figure 1) were taken with a rosette sampler equipped with

twelve 10–litre PVC Niskin bottles with stainless–steel internal springs,

while water samples from stns 01 and 03, located at the Northern and

Southern ends of the segment, were taken with 5-litre PVC Niskin bottles.

Page 226: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

208

9.0º 8.9º 8.8º 8.7ºLongitude (West)

42.1º

42.2º

42.3º

Latit

ude

(Nor

th)

Rande Strait San

Sim

on B

ay

C. MarCíes Islands

1000 1500 2000

-40

-30

-20

-10

07.3-10.8NS CS

5.9-15.2SS

6.3-10.6

NB SB6.2-15.3

N(R03) S(R01)

9.5-17.0

P.Borneira

DCM12 & CTD st.

CTD station

Vigo

Pontevedra

Arousa

River Oitabén Verdugo

Silleiro buoy

Rande Strait

Iberian Peninsula

2

3

4

5

9

1

00

Ría de VigoR03

R01

b

a

c

At stn 00, salinity and temperature were recorded with a SBE 9/11

conductivity–temperature–depth probe attached to the rosette sampler.

Conductivity measurements were converted into practical salinity scale

values with the equation of UNESCO (1985). Water samples were

collected from five depths: the surface (50% Photosynthesis Available

Radiation, PAR; 2.5±0.3 m), the depth of the 25% PAR (7±1), the depth of

the 1% PAR (14.2±2.2 m), 26.2±1.2m and the bottom (40.3±0.5 m).

Figure 1. Chart of the Ría de Vigo (NW Spain), indicating the sampling sites: stns 00 (where the ADCP current meter was moored), 01 and 03. The five sectors ―three at the surface (NS, CS and SS) and two at the bottom (NB and SB) ― defined to study the lateral variability of the middle ría (V–shaped, 2.5 km wide, 40 m maximum depth) are presented in inset a. Inset b shows the longitudinal transect repeated during the 60 visits to the Ría de Vigo performed from 1990 to 1997 (see text for details). The shaded zone in inset b is the free surface area of the volume bounded by the study transverse section. Inset c shows the Ría de Vigo in relation to the Rías Baixas and the location of the Seawatch buoy of Puertos del Estado off Cape Silleiro.

Page 227: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

209

At stns 01 and 03, salinity and temperature were recorded with a CTD

SBE-25 and water samples were collected from two depths: surface and

bottom.

Wind and current measurements. Coastal winds were taken from the

Seawatch buoy off Cape Silleiro (http://www.puertos.es/; Figure 1), a

representative site for winds blowing off the Ría de Vigo (Herrera et al.

2005). Low–pass filtered winds are used to define the prevailing

oceanographic scenario in the ría during each period. The heat balance

and freshwater flux for each sampling date were taken from Piedracoba et

al. (2005a).

Current velocity profiles were measured aboard R/V Mytilus with a

RDI Vessel Mounted Broad Band Acoustic Doppler Current Profiler

(VMADCP) of 300 kHz along the transverse section joining stns 01 and 03

on 18 and 22 July and 19 September 2002. In addition, current profiles

were measured with an Aandera DCM12 Acoustic Doppler Current Meter

(ADCP) moored at stn 00. The mooring was deployed from 10 to 29 July

and from 17 to 26 September 2002. The DCM12 lied on the seabed,

recording at 30 minutes intervals the current integrated in 5 vertical cells

11 m thick. DCM12 and VMADCP data were rotated counter clockwise

30º to produce a longitudinal and a transversal component referred to the

main axis of the ría.

A C++ version of the HAMSOM model, recently validated in the outer

(Souto et al. 2003) and inner (Diz et al. 2004) Ría de Vigo, was used to

reproduce the lateral variability of the subtidal current fields on 15 and 26

July and 17, 23 and 26 September 2002. The accuracy of the model to

reproduce the lateral current field has been successfully tested in

Piedracoba et al. (submitted, Chapter 3).

Page 228: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

210

Horizontal advection. Five sectors ―three at the surface (NS, CS and

SS) and two at the bottom (NB and SB) ― were defined to account for the

lateral variability in the middle ría (Figure 1). The criterion to divide the

surface and bottom sectors was based on the level of no–horizontal

motion, considering the current profiles measured with the DCM12 and

the VMADCP, together with the outputs of the HAMSOM model.

Horizontal advection velocities were averaged in the 5 sectors.

Vertical advection. The Ría de Vigo was visited with mass balance

purposes 60 times from 1990 to 1997. The inset of Figure 1 shows the

longitudinal transect repeated during the 60 visits. The good relationship

between the bottom horizontal flow (QB) and the vertical flow (QZ)

obtained from the mass balances of the 60 visits (Figure 1), allows

calculating the unknown QZ ( in m3 s-1) from the measured QB ( in m3 s-1)

during the sampling dates in July and September 2002:

BZ Q0.03)0.73(43)212(Q ⋅±+±−= r = +0.92, p < 0.001 (1)

The vertical convective velocity (VZ) was calculated considering the

surface area of the shaded zone in the inset of Figure 1. The vertical

displacement of the LNM between two consecutive surveys was

calculated too.

Vertical mixing. The turbulent mixing coefficient, Kz (m2 d-1), was

parameterized according to Munk and Anderson (1948), using the ADCP

and CTD data.

21

i0 )R10(1KKz−

⋅+⋅= (2)

ZU103K 03

0 ⋅⋅⋅= − (3)

Page 229: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

211

2i

zu

zg

R

⎟⎠⎞

⎜⎝⎛∂∂

∂ρ∂

⋅ρ

−= (4)

Where 0U is the amplitude of the tidal velocity from 0 to Z m, with Z

been the water column depth; iR is the Richardson number; and ρ is the

average density. Once Kz was obtained, velocity dimensions were

obtained dividing Kz by Z/2, the distance between the gravity centres of

the upper and lower layer.

Stability. The water column average Brunt-Väsisälä frequency, <N> (in

mín-1) at stn 00 (Figure 1) was calculated as <N>

= [ ]{ } 21ulul Zg4 /(/)( ρ+ρρ−ρ , where ρl and ρu are the average density of

the upper and lower layers and g is the gravity acceleration.

Chemical variables. Salinity (S, accuracy ± 0.003), N-nutrients (NH4+, ±

0.05 µM; NO2–, ± 0.02 µM; and NO3–, ± 0.1 µM), dissolved organic nitrogen

(DON, ± 0.2 µmol N L-1), particulate organic nitrogen (PON, ± 0.1 µmol N

L-1) and chlorophyll (Chl a, ± 0.05 mg m-3) have been measured as

described elsewhere (Álvarez-Salgado et al. 2005, Chapter 4).

Horizontal/vertical fluxes. Average concentrations (µmol Kg-1) of all

species were calculated from the three stations located along the middle

segment of the ría by means of an inverse distance interpolation method.

The horizontal flux of any species (Fx, in µmol m-2 d-1) was calculated as

the product of horizontal velocity ( XV ) times concentration and times

density in the 5 sectors defined above. Vertical convective fluxes ( ZF , in

µmol m-2 d-1) were obtained multiplying the vertical convective velocities

( ZV ) calculated above times the depth-average concentration in the study

Page 230: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

212

section and times density. Finally, vertical diffusive fluxes ( ZM , in µmol

m-2 d-1), were obtained as:

)( SBz

Z CCz

KM −∆

= (5)

Where CS and CB are the average concentrations of any species in the

surface and bottom layer, respectively.

Geochemical budget. The net budget of inputs minus outputs, i – o ,

of any species (C) into a given volume results from the accumulation,

t/)VC( ∆⋅∆ , and net ecosystem production, NEPC, terms:

CNEPtVCoi −

∆⋅∆

=−)( (6)

The accumulation term can be separated in two terms:

tVC

tCV

tVC

∆∆⋅+

∆∆⋅=

∆⋅∆ )( (7)

The first term on the right side of eq. (7) indicates the net accumulation

due to changes in concentration and the second term due to changes in

volume. The average NEPC between two consecutive surveys can be

estimated as:

tCVioNEPC ∆⋅∆

+−=)( (8)

Where the balance of o - i (in mol C m-2 d-1) is calculated as:

AAFAFio BBSS /)( ⋅−⋅=− (9)

FS and FB are the average surface and bottom horizontal fluxes (in µmol

m-2 d-1); AS and AB are the areas of the surface and bottom layer of the

transverse section (in m2); and A is the free surface area of the ría bounded

by the transverse section (35.6 x 106 m2).

Page 231: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

213

3. Results

3.1. Water flows

July 2002

Three periods can be defined from Figures 2a, d. The first, from 15 to 18

July, was characterized by an intense upwelling episode that reached the

maximum intensity on 15 July. A strong surface outflow of –12 cm s-1 was

recorded. During the second period, from 18 to 22 July, shelf northerly

winds relaxed (they ranged from 0 to –3 m s-1), producing a reversal of the

residual circulation pattern (surface inflow of < 7 cm s-1, and bottom

outflow over –3 cm s-1). Finally the third period, from 22 to 26 July, was

dominated by northerly winds that promoted an intense upwelling on 26

July. A positive residual circulation pattern established again, with surface

and bottom currents similar to the first period.

Temperature rather than salinity gradients controlled the density

distribution in the middle ría (Figures 2e, f). Continental runoff was very

low throughout the study period (<9 m3 s-1; Figure 2b). On 15 July, the

salinity and temperature distributions showed the typical response to an

upwelling event. Despite the subsequent relaxation of shelf winds, the

effect of coastal upwelling still remained in the bottom layer 3 days later,

when higher salinities and lower temperatures were recorded. The

reversal of the circulation in response to the wind calm (Figure 2a) had a

clear impact on the water column: salinity was more homogenous and the

halocline was difficult to distinguish while surface temperature increased

and the thermocline was found at 20 m depth. The reduced river

discharge left the estuary by the surface layer, and saltier and colder water

entered into the ría through the bottom layer.

Page 232: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

214

The analysis of the longitudinal component of the residual current in

the 5 sectors shows a positive 2–layered residual circulation pattern under

upwelling conditions (Figures 3a, d). Other circulation patterns consisted

of 2 layers in the central and southern sectors and a net inflow through the

northern sector during the summer upwelling relaxation (Figure 3b) and a

transient net entry of water during upwelling spin down (Figure 3c).

Vertical convective velocities were upwards during upwelling events, 3.7

± 0.5 m d-1 and 4.6 ± 1.4 m d-1 on 15 and 26 July respectively, whereas they

were nearly zero during the relaxation of upwelling (18 July)and reversed

to downwards on 22 July (0.0 ± 0.5 m d-1 and -1.0 ± 0.1 m d-1 respectively).

Mixing velocities reached a maximum during upwelling events, being

even twice the vertical convective velocity on 15 July (8.2 ± 0.2 m d-1) and

half (2.4 ± 0.1 m d-1) during the other intense upwelling (26 July), as a

result of the stratification that took place during the relaxation between

the two upwelling episodes. However, during the upwelling relaxation

(from 18 to 22 July) mixing ranged between 0 and 2 m d-1. The vertical

displacement of the LNM was negligible compared with the vertical

convective and mixing velocities.

The four sampling dates in July will be studied grouped in the

following three intervals: i) spin down of strong summer upwelling (15-18

Jul), ii) summer relaxation (18-22 Jul) and iii) spin up of weak summer

upwelling (22-26 Jul).

Page 233: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

215

Figure 2. Time evolution of shelf winds, in m s–1 (a); continental runoff (Qr), precipitation (P) and evaporation (E), in m3 s–1 (b); heat balance, in cal cm–2 d–1 (c); residual currents, in cm s–1 (d);salinity (e); temperature, in °C (f); dissolved organic nitrogen, in µmol L–1 (g), ammonium, in µmol kg–1 (h); chlorophyll a, in µg L–1 (i); particulate organic nitrogen, in µmol L–1 (j), nitrate, in µmol kg–1 (k), and nitrite, in µmol kg–1 (l) during July 2002. Heat and freshwater fluxes taken from Piedracoba et al. (2005a).

-12-606

12

16/7 18/7 20/7 22/7 24/7 26/7-202468

101214

Qr-

P-E

Qr P E

-200-100

0100200300400

H

16/7 18/7 20/7 22/7 24/7 26/7 16/7 18/7 20/7 22/7 24/7 26/7

a b c

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

d

j

g i

-10

-20

-30

e f

h

k

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

15/7

18/7

22/7

26/7

15/7

18/7

22/7

26/7

15/7

18/7

22/7

26/7

l

Page 234: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

216

1.7±0.1-0.6±0.5

-3.0±0.1

17/09/02

-0.6±0.2 0.2±0.0 0.6±0.2

0.5±0.1 0.1±0.0

3.8±0.10.8±0.3

2.2±0.4

23/09/02

-0.8±0.2 -0.7±0.2 -0.7±0.2

-0.7±0.2 -0.5±0.1

2.1±0.10.0±0.3

1.2±0.2

26/09/02

-1.1±0.3 -1.0±0.3 0.3±0.1

0.3±0.1 0.5±0.1

19/09/02

1.1±1.0 1.8±0.8 4.1±1.2

-4.8±0.9 -4.4±0.7

5.3±0.20.5±0.6-3.7±0.1

N S N S

Figure 3. Time course of horizontal (km d-1) and vertical velocities (m d-1) for each sampling date. Solid circle, outflow (V < 0 or westwards) and crossed circle, inflow (V > 0 or eastwards) into the ría; solid arrow, vertical advective velocity (upward, >0; downward, <0); curved arrow, vertical mixing velocity (Kz/∆z); white arrow, vertical displacement of the LNM.

1.7±1.3 -2.9±1.5 -7.0±1.6

2.9±0.4 3.4±0.3

18/07/02

0.7±0.1-0.1±0.30.0±0.5

22/07/02

9.7±1.6 9.3±0.5 7.3±1.8

8.4±0.9 6.6±0.6

1.5±0.10.1±0.3-1.0±0.1

8.2±0.20.0±0.3

3.7±0.5

15/07/02

-4.5±1.4 -4.0±1.2 -3.7±1.1

2.3±0.7 1.7±0.5

2.4±0.10.4±0.1

4.6±1.4

26/07/02

-1.3±0.4 -2.1±0.6 -3.0±0.9

-0.5±0.1 2.1±0.6

N S N S

Page 235: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

217

Figure 4. Time evolution of shelf winds, in m s–1 (a); continental runoff (Qr), precipitation (P) and evaporation (E), in m3 s–1 (b); heat balance, in cal cm–2 d–1 (c); residual currents, in cm s–1 (d);salinity (e); temperature, in °C (f); dissolved organic nitrogen, in µmol L–1 (g), ammonium, in µmol kg–1 (h); chlorophyll a, in µg L–1 (i); particulate organic nitrogen, in µmol L–1 (j), nitrate, in µmol kg–1 (k), and nitrite, in µmol kg–1 (l). Heat and freshwater fluxes taken from Piedracoba et al. (2005a).

-12-606

12

17/9 19/9 21/9 23/9 25/9-202468

101214

Qr-

P-E

Qr P E

-200-100

0100200300400

H

16/9 18/9 20/9 22/9 24/9 26/9 16/9 18/9 20/9 22/9 24/9 26/9

-40

-30

-20

-10

-40

-30

-20

-10

d fe

g h i

j lk

-10

-20

-30

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

17/9

19/9

23/9

26/9

17/9

19/9

23/9

26/9

17/9

19/9

23/9

26/9

-40

-30

-20

-10

a b c

Page 236: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

218

September 2002

Three periods can also be defined during September on basis of Figures

4a, d. The first period, from 17 to 19 September, was dominated by

southeasterly winds that decreased gradually from 6 to 3 m s-1. They

originated a negative residual current pattern with a maximum surface

inflow of 10 cm s-1 and bottom outflow of –10 cm s-1. The second period,

from 19 to 23 September, was characterized by the transition from

southerly to northerly winds, when the reversal of the residual pattern

persisted but current velocities diminished. On 23 September, the

circulation evolved to 3-layes, with outflow through the surface and

bottom layers and inflow through the intermediate layer. Finally, the third

period, from 23 to 26 September, was characterized by moderate

northeasterly winds of less than –4 m s-1 blowing over the shelf and the 2-

layers positive circulation pattern was re-established (surface outflow of

about –5 cm s-1 and bottom inflow of 5 cm s-1). The freshwater flow was

again very limited: from 5.7 to 14.5 m3 s-1 (Figure 4c).

The thermohaline structure found on 17 September corresponded with

a typical autumn downwelling, with warm (>17 ºC) and relatively salty

(>34.8) shelf surface waters occupying the surface layer of the middle ría

(Figures 4e, f). The persistence of shelf southerly winds produced a

deepening of the surface salinity and temperature as a consequence of

downwelling. Downwelling contributed to homogenize salinity through

the water column, but the solar irradiation was able to keep the

temperature gradient. Southerly winds changed to northeasterly on 22

September, and this effect was recorded in the water column with a slight

uplift of the isohalines and isotherms, suggesting the entry of oceanic

ENACW through the bottom layers on 26 September.

Page 237: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

219

From the analyses of the longitudinal component of the residual

current in the 5 sectors, it was distinguished a first period when the

circulation pattern consisted of a net entry with low velocities (Figure 3e)

and practically there was no horizontal movement in the ría, and a

downwelling episode characterized by a negative circulation pattern

(Figure 3f). After the transition period from southerly to northerly winds,

the ría responded with a net outflow with similar magnitude to those

recorded on 17 September (Figure 3g). Finally, moderate upwelling took

place in the ría (Figure 3h). Maximum velocities were recorded on 19

September, when the ría showed the maximum surface asymmetry. The

vertical convective velocities changed from downwards on 17 and 19

September, -3.0 ± 0.1 m d-1 and -3.7 ± 0.1 m d-1 respectively, to upwards on

23 and 26 September, 2.2 ± 0.4 m d-1 and 1.2 ± 0.2 m d-1 respectively. The

mixing velocity was always higher than the convective vertical velocity,

except on 17 September, and it reached the maximum on 19 September

(5.3 ± 0.2 m d-1). Again, the vertical displacements of the LNM were

negligible compared with the vertical convective and mixing velocities.

As in July, the four sampling dates in September were grouped in three

intervals according to the hydrodynamic conditions: i) autumn

downwelling (17-19 Sep), ii) autumn transition (19-23 Sep) and iii) autumn

upwelling (23-26 Sep).

Page 238: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

220

Figure 5a. Averaged concentrations of NT, DON and PON in µmol kg–1 and contributions of NO3 -, NO2- NH4+ to NT for each sector and July survey date.

15/07/02

18/07/02

15/07/02

18/07/02

22/07/02 22/07/02

68.5±3.45.0±0.326.5±1.5

73.1±2.52.9±0.224.0±1.0

78.1±6.93.1±0.4

18.8±2.3

67.1±0.92.0±0.1

30.9±0.4

67.3±1.02.2±0.130.5±0.5NO3

-

NO2-

NH4+

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

3.5±0.25.1±0.14.1±0.2

4.9±0.25.9±0.23.8±0.2

1.9±0.28.7±0.24.5±0.2

13.4±0.24.5±0.12.1±0.2

12.2±0.24.5±0.12.5±0.2

NTDONPON

8.5±0.24.6±0.13.7±0.2

7.8±0.24.4±0.14.6±0.2

4.4±0.25.0±0.34.5±0.2

14.5±0.23.9±0.21.2±0.2

14.0±0.23.9±0.21.4±0.2

NTDONPON

0.9±0.25.9±0.23.7±0.2

3.5±0.26.1±0.24.1±0.2

1.0±0.26.3±0.13.9±0.2

13.7±0.24.2±0.12.8±0.2

12.1±0.24.3±0.12.8±0.2

NTDONPON

73.5±1.42.1±0.124.4±0.6

69.3±1.52.2±0.1

28.5±0.7

60.9±2.52.8±0.2

36.3±1.4

75.0±0.91.6±0.1

23.4±0.3

75.1±0.91.8±0.123.1±0.3

28.3±9.12.9±0.9

68.8±10.9

63.1±3.23.1±0.2

33.8±1.7

32.0±8.63.1±0.8

64..9±9.4

68.2±0.92.0±0.129.8±0.4

67.9±1.02.0±0.130.1±0.4

% of NT

% of NT

% of NT

26/07/02 26/07/02

NO3-

NO2-

NH4+

8.2±0.25.1±0.13.6±0.3

7.7±0.25.1±0.13.8±0.5

7.1±0.25.0±0.12.9±0.3

12.9±0.24.7±0.11.6±0.5

12.7±0.24.8±0.11.8±0.4

NTDONPON

65.3±1.42.7±0.132.0±0.7

71.9±1.63.1±0.1

25.0±0.6

69.3±1.72.3±0.1

28.4±0.7

79.0±1.02.4±0.1

18.7±0.3

78.2±1.02.4±0.1

19.4±0.3

% of NT

Page 239: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

221

Figure 5b. Averaged concentrations of NT, DON and PON in µmol kg–1 and contributions of NO3 -, NO2- NH4+ to NT for each sector and September survey date.

17/09/02

19/09/02

17/09/02

19/09/02

23/09/02 23/09/02

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NTDONPON

NTDONPON

% of NT

% of NT

% of NT

2.8±0.25.9±0.13.8±0.2

1.6±0.28.0±0.12.9±0.2

0.8±0.26.2±0.24.0±0.2

6.2±0.27.5±0.11.6±0.2

6.2±0.27.2±0.11.9±0.2

2.0±0.26.9±0.15.6±0.2

2.4±0.25.4±0.24.2±0.3

2.2±0.24.8±0.13.9±0.2

4.9±0.25.2±0.11.3±0.2

4.5±0.25.6±0.11.9±0.2

1.3±0.26.3±0.16.0±0.2

1.6±0.26.4±0.15.6±0.2

1.8±0.26.4±0.15.5±0.2

5.8±0.26.7±0.12.8±0.2

5.3±0.26.6±0.13.4±0.2

23.0±2.63.2±0.373.7±3.5

20.8±4.44.7±0.674.5±6.2

13.7±8.15.6±1.3

80.7±13.4

37.5±1.46.1±0.256.4±1.3

37.2±1.45.9±0.256.9±1.3

21.2±3.56.3±0.6

72.5±4.8

25.1±3.15.1±0.469.8±3.9

15.8±3.04.2±0.4

80.0±4.8

16.5±1.43.8±0.279.7±2.1

16.7±1.54.0±0.2

79.3±2.3

30.6±6.16.7±0.9

62.7±6.6

24.9±4.87.2±0.867.9±5.9

17.0±3.65.3±0.6

77.7±5.6

23.4±1.34.6±0.272.0±1.7

23.4±1.44.6±0.272.0±1.8

26/09/02 26/09/02

NO3-

NO2-

NH4+ % of NT

4.0±0.27.4±0.13.8±0.2

3.7±0.27.1±0.12.9±0.2

3.3±0.26.3±0.13.5±0.2

5.9±0.28.0±0.13.0±0.2

5.9±0.27.8±0.13.1±0.2

26.4±1.94.9±0.2

68.6±2.2

26.0±2.05.3±0.368.8±2.5

24.0±2.25.1±0.3

70.9±2.9

25.1±1.34.8±0.270.1±1.6

25.3±1.34.8±0.269.9±1.6

Page 240: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

222

Figure 6a. Averaged horizontal convective fluxes of NT, DON, PON, NO3 -, NO2- and NH4+ (Fx, µmol m-2 d-1, <0 outflow, >0 inflow) for each sector and July survey date.

15/07/02

18/07/02

15/07/02

18/07/02

22/07/02 22/07/02

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NTDONPON

NTDONPON

-13±2-19±3-15±2

-20±3-24±3-15±2

-8±1-40±5-21±3

31±510±25±1

21±38±14±1

-60±6-32±4-26±3

-22±5-12±3-13±3

8±39±58±3

43±312±14±1

47±213±15±1

6±243±628±4

33±257±238±2

9±261±638±4

115±735±224±2

80±429±219±1

-8.8±1.1-0.6±0.1-3.4±0.4

-14.4±2.0-0.6±0.1-4.8±0.6

-6.6±0.8-0.3±0.1-1.6±0.1

20.5±3.20.6±0.19.5±1.5

14.0±2.20.5±0.16.4±1.0

-43.9±4.8-1.2±0.1

-14.6±1.5

-15.5±3.8-0.5±0.1-6.3±1.6

4.6±1.80.2±0.12.7±1.1

31.9±2.50.7±0.110.0±0.8

35.6±1.50.8±0.110.9±0.5

1.8±0.60.2±0.14.4±0.7

20.8±1.01.0±0.1

11.1±0.5

3.0±0.70.3±0.16.0±0.8

78.1±4.42.3±0.234.1±2.0

54.2±2.82.3±0.224.0±1.3

26/07/02 26/07/02

NO3-

NO2-

NH4+

-24±3-15±2-11±1

-16±2-10±1-8±1

-9±1-6±1-4±1

-6±1-2±1-1±1

27±410±24±1

-15.8±2.2-0.7±0.1-7.7±1.1

-11.5±1.6-0.5±0.1-4.0±0.5

-6.2±0.9-0.2±0.1-2.5±0.3

-4.7±0.7-0.1±0.0-1.1±0.2

20.9±3.20.6±0.15.2±0.8

NTDONPON

NTDONPON

Page 241: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

223

Figure 6b. Averaged horizontal convective fluxes of NT, DON, PON, NO3 -, NO2- and NH4+ (Fx, µmol m-2 d-1, <0 outflow, >0 inflow) for each sector and September survey date.

17/09/02

19/09/02

17/09/02

19/09/02

23/09/02 23/09/02

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NO3-

NO2-

NH4+

NTDONPON

NTDONPON

NTDONPON

8.3±1.627.9±4.422..9±3.7

4.4±1.19.9±2.37.5±1.8

2.1±1.14.6±2.43.7±1.9

-23.1±1.6-25±2.1-6.2±0.1

-19.7±1.1-24.4±1.7-8.1±0.2

-1.0±0.1-4.5±0.7-4.3±0.6

-1.2±0.1-4.7±0.7-4.2±0.6

-1.5±0.1-5.1±0.7-4.4±0.6

-4.1±0.5-4.7±0.7-1.9±0.2

-2.6±0.3-3.3±0.5-1.7±0.2

0.4±0.10.06±0.01

1.3±0.2

0.05±0.020.01±0.0030.19±0.03

-0.1±0.02-0.03±0.001-0.38±0.04

1.2±0.20.19±0.031.7±0.3

0.3±0.050.05±0.010.5±0.1

1.8±0.50.5±0.1

6±1

1.1±0.30.2±0.13.1±0.7

0.3±0.20.1±0.051.6±0.8

-3.8±0.1-0.9±0.1-18.4±1.6

-3.3±0.04-0.8±0.04-15.6±1.1

1.7±0.33.6±0.62.3±0.4

0.3±0.11.3±0.20.5±0.1

-0.5±0.01-3.7±0.5-2.4±0.3

3.1±0.53.7±0.60.8±0.2

0.8±0.10.9±0.10.2±0.05

-0.3±0.01-0.1±0.01-0.6±0.1

-0.3±0.01-0.08±0.01-0.8±0.1

-0.2±.0.01-0.1±0.01-1.1±0.1

-0.9±0.1-0.2±0.02-2.9±0.4

-0.6±0.1-0.1±0.02-1.9±0.3

26/09/02 26/09/02

NO3-

NO2-

NH4+

NTDONPON

1.0±0.21.9±0.31.0±0.2

-3.7±0.5-7.0±1.0-3.8±0.5

-3.6±0.4-6.9±1.0-3.8±0.5

1.8±0.32.4±0.40.9±0.2

2.8±0.53.8±0.61.5±0.3

0.3±0.050.05±0.010.7±0.1

-1±0.1-0.2±0.02-2.6±0.4

-0.9±0.1-0.2±0.02-2.6±0.4

0.4±0.10.08±0.011.2±0.2

0.7±0.10.13±0.022.0±0.3

Page 242: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

224

3.2. Nitrogen fluxes

July 2002

NO3- concentrations (Figure 2k) were maximum at the bottom layer

during the spin down of the strong summer upwelling (15-18 July)

because of the entry of ENACW. In fact, NO3- was highly correlated with

temperature (r2= 0.92, n= 20, p< 0.0001). On the contrary, NH4+ and NO2-

(Figure 2h, l) cannot be explained only by temperature for the same period

(r2= 0.69, n= 20, p< 0.0001 and r2= 0.63, n= 20, p< 0.0001, respectively). The

levels of NO2- were maximum in the bottom layer but <0.4 µmol Kg-1. In

the surface layer, maximum concentrations of Chl a (Figure 2i) and PON

(Figure 2j) were recorded. During the summer relaxation (18-22 July),

lower Chl a and PON were recorded in the surface compared with the

previous period and a subsurface Chl a maximum developed at 20 m

depth (the nutricline). DON profiles (Figure 2g) followed the variability of

temperature (r2= 0.69, n= 20, p< 0.0001), and concentrations ~4 µmol Kg-1

were recorded for temperatures <12ºC. DON accumulation in the surface

layer took place at the end of the summer relaxation, reaching

concentrations >7 µmol Kg-1. Finally, during the spin up of weak summer

upwelling (22-26 July), nutrient-rich water entered through the bottom

layers, but this entry was weaker compared with the first upwelling

period (bottom NO3- < 11 µmol Kg-1).

No significant differences in either nutrient stocks or fluxes were

observed between the two bottom sectors. On the contrary, significant

differences, specially regarding nutrient fluxes, were observed between

the three surface sectors (Tables 1a, 1b). The spatial differences were

significant for NO3- (Figures 4, 5a). DON and PON stocks were very

homogenous through the surface and bottom layers, whereas the fluxes

Page 243: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

225

varied without a defined pattern. On July 15 maximum DON and PON

surface outflows occurred in the northern sector. However, on July 18 and

26 maximum surface outflows occurred in the southern sector.

NO3- was the most abundant N–nutrient, representing 69% of NT, (=

NO3- + NO2- + NH4+), followed by NH4+, 29%, and NO2-, 2%. NO3- and

NO2- were more abundant at the lower layer and maximum NH4+ was

normally found at the surface layer (Figure 5a). The highest surface NT

concentrations were found in the central and southern sectors, while in

the bottom the concentration was higher in the northern sector. DON was

always higher in the surface sectors.

During the strong summer upwelling (15-18 Jul), NT fluxes intensified

through the surface southern sector and the bottom entry of nutrients

presented significant differences along the section at the beginning (15 Jul)

and no differences at the end (18 Jul) of the strong summer upwelling

(Figure 6). During the summer relaxation (22 Jul), the net entry of NT was

larger through the northern bottom sector and, finally during the spin up

of weak summer upwelling differences in the bottom sectors were found

(inflow by the southern and outflow by the northern sector). DON and

PON outgoing fluxes were larger in the southern surface sector under

strong upwelling and weak upwelling but during the relaxation the

largest ingoing DON and PON fluxes were recorded in the northern

sector.

September 2002

The most abundant N-nutrient during September was NH4+ (Figures

4h, k). Only 53% of the variability of NH4+ was explained by temperature

(r2= 0.53, n= 20, p< 0.0001). NO3- and NO2- also correlated with

temperature (r2= 0.72, n= 20, p< 0.0001, r2= 0.69, n= 20, p< 0.0001,

Page 244: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

226

respectively). During the autumn downwelling (17–19 September), there

was an entry of nutrient-poor waters from the shelf. Concomitantly, lower

PON (~4 µmol Kg-1) and Chl a (~4 µmol Kg-1) were measured at the

surface layer compared with July (Figure 4i, j). DON (Figure 4j) was

higher than PON during this period and presented a homogenous

distribution throughout the water column. At the beginning of the

autumn transition (19-23 September), nutrients levels were still very low

in the surface layer, but the wind change caused a progressive entry of

ENACW in the bottom layer. The levels of NO3- and NH4+ (Figures 4h, k)

increased specially at the beginning of the autumn upwelling (23-26

September), although the concentration of NO3- at the bottom was five

times less than the levels measured during the weakest summer

upwelling. The levels of NO2- were higher in the bottom layer where a

maximum of 0.8 µmol Kg-1 was recorded during the autumn

downwelling. Accordingly, high surface Chl a (> 8 µmol Kg-1) and PON (>

7 µmol Kg-1) were measured (Figure 4i, j). As for July, the DON maximum

(Figure 4g) was delayed 3-4 days compared with the Chl a and PON

maxima.

In September, most of the concentration differences along the

transverse section occurred in organic nitrogen. As in July, the fluxes of

inorganic and organic species were significantly different between sectors

(Tables 2a, 2b).

The NT concentration reduced to more than a half compared with July

and the partitioning was: 23%, 5% and 72% for NO3-, NO2- and NH4+

respectively. The most abundant nutrient (NH4+) presented the highest

concentration at the lower layer during the autumn downwelling (19 Sept)

and the autumn transition (23 Sept) and a nearly homogenous distribution

Page 245: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

227

in the water column during the weak upwelling of 26 September (Figure

5b). During this period, the distribution of NT was nearly homogenous in

the surface and bottom sectors.

DON was slightly higher in September than in July. DON was either

higher in the bottom sectors or was homogenously distributed in water

column due to the dominant downward vertical fluxes. This distribution

was linked to the dominant residual circulation pattern in each period.

Under upwelling conditions, the upward vertical advection that supplies

nutrients to the photic layer promotes a high primary production

(Piedracoba et al. this issue) and therefore, high concentrations of organic

matter in the upper layer. However, the low nutrient fluxes under

downwelling conditions are the reason behind the low organic matter

production compared with upwelling conditions.

Nutrient fluxes in September were up to an order of magnitude lower

than in July. The inflow of inorganic/organic N species was larger

through the southern sector and the outflow through the bottom layer was

specially intensified on 19 September. Although a net outflow of

inorganic/organic N species was recorded in the weak autumn upwelling

(23-26 Sep), the fluxes were extremely low compared with summer

upwelling (Figure 6b).

Page 246: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

228

concentrations

Date NT DON PON NO3- NO2- NH4+ 15/07/2002 SS-CS x x n.s n.s x n.s CS-NS x x x n.s n.s x SS-NS x x n.s n.s x x NB-SB x n.s n.s n.s n.s n.s 18/07/2002 SS-CS x n.s x x n.s x CS-NS x x n.s x x x SS-NS x n.s x x x x NB-SB x n.s n.s n.s n.s n.s 22/07/2002 SS-CS x n.s n.s x n.s x CS-NS x n.s n.s x n.s x SS-NS n.s x n.s n.s n.s n.s NB-SB x n.s n.s n.s n.s n.s

26/07/2002 SS-CS x n.s n.s x x x

CS-NS x n.s x n.s x x SS-NS x n.s x x x x NB-SB n.s n.s n.s n.s n.s x

Table 1a. t-TEST for the concentrations of NT, DON, PON, NO3 -, NO2- and NH4+ between the three surface (NS, CS and SS) and the two bottom (NB and SB) sectors, for each sampling date in July 2002; n.s: non-significant differences; * : p <0.05;

Page 247: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

229

fluxes

Date NT DON PON NO3- NO2- NH4+ 15/07/2002 SS-CS x n.s n.s x n.s x CS-NS x x x x n.s x SS-NS x x x x x x NB-SB x n.s n.s x n.s x 18/07/2002 SS-CS x x x x x x CS-NS x x x x x x SS-NS x x x x x x NB-SB n.s n.s n.s n.s n.s n.s 22/07/2002 SS-CS x x x x x x CS-NS x n.s n.s x x x SS-NS n.s x x n.s n.s x NB-SB x x x x n.s x

26/07/2002 SS-CS x x x n.s n.s x

CS-NS x x x x x x SS-NS x x x x x x NB-SB x x x x x x

Table 1b. t-TEST for the horizontal convective fluxes of NT, DON, PON, NO3 -, NO2- and NH4+ between the three surface (NS, CS and SS) and the two bottom (NB and SB) sectors, for each sampling date in July 2002; n.s: non-significant differences; * : p <0.05;

Page 248: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

230

concentrations

Date NT DON PON NO3- NO2- NH4+ 17/09/2002 SS-CS x x x n.s x n.s CS-NS x x x n.s n.s n.s SS-NS x n.s n.s n.s x n.s NB-SB n.s x n.s n.s n.s n.s 19/09/2002 SS-CS x x x n.s x n.s CS-NS n.s x n.s x x x SS-NS n.s x x n.s x n.s NB-SB n.s x x n.s n.s n.s 23/09/2002 SS-CS n.s n.s n.s n.s n.s n.s CS-NS n.s n.s n.s n.s x n.s SS-NS x n.s x x n.s n.s NB-SB x n.s x n.s n.s n.s 26/09/2002 SS-CS n.s x x n.s n.s n.s

CS-NS n.s x x x n.s n.s SS-NS x x n.s n.s n.s n.s NB-SB n.s n.s n.s n.s n.s n.s

Table 2a. t-TEST for concentrations of NT, DON, PON, NO3 -, NO2- and NH4+ between the three surface (NS, CS and SS) and the two bottom (NB and SB) sectors, for each sampling date in September 2002; n.s: non-significant differences; * : p <0.05;

Page 249: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

231

fluxes

Date NT DON PON NO3- NO2- NH4+ 17/09/2002 SS-CS x x x x x x CS-NS x x x x x x SS-NS x x x x x x NB-SB x x x x x x 19/09/2002 SS-CS x x x n.s x x CS-NS x x x x n.s x SS-NS x x x x x x NB-SB x n.s x x n.s x 23/09/2002 SS-CS n.s n.s n.s n.s n.s n.s CS-NS x n.s n.s x n.s x SS-NS x n.s n.s x n.s x NB-SB x x n.s x x x 26/09/2002 SS-CS x x x x x x

CS-NS n.s n.s n.s n.s n.s n.s SS-NS x x x x x x NB-SB x x x x x x

Table 2b. t-TEST for horizontal convective fluxes of NT, DON, PON, NO3 -, NO2- and NH4+ between the three surface (NS, CS and SS) and the two bottom (NB and SB) sectors, for each sampling date in September 2002; n.s: non-significant differences; * : p <0.05;

Page 250: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

232

3.3. Nitrogen budget

July 2002

A budget of o – i (in mg N m-2 d-1) for NT, PON and DON was built up

considering the fluxes per m2 of free surface area from the middle section

to the inner reaches of the ría (Table 3). In addition, the accumulation

terms of all nitrogen species were also calculated (V·dN/dt), considering

that the concentrations of NT, PON and DON in the volume of the ría

bounded by the middle section do not differ significantly from the

concentrations in the middle section.

NEP can be obtained summing up o – i and V·dN/dt (see eq. 8) for NT

(NEPT), DON (NEPD) and PON (NEPP) between two consecutive surveys.

During the strong summer upwelling, the volume of the ría bounded by

the middle section trapped -289 ± 17 mg N m-2 d-1, i.e. 62% of the NT entry,

but only 44% was transformed into organic matter (NEPT = -127 ± 37 mg N

m-2 d-1), while 162 mg N m-2 d-1 accumulated into the volume. On the

contrary, the ría was a nutrient source during the subsequent upwelling

relaxation (o – i = 101 ± 5 mg N m-2 d-1) at the expenses of the nutrients

accumulated during the previous period. Finally, during the following

weak upwelling episode, nutrient salts were trapped again, 57% of them

being destined to produce organic matter and 43 % to accumulate into the

study volume. During July, the % of NEPT due to NO3- and NH4* was 67%

and 33% respectively, in average.

The fraction of NEPT transferred to the sediments (NEPSED) can be

inferred indirectly from the following mass balance equation:

δ0NEPNEPNEPNEP SEDPDT ±=+++ (10)

where, NEPSED represents the balance between PON sedimentation and

PON resuspension + NT diffusion from the sediment (Álvarez–Salgado et

Page 251: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

233

al. 1996a). It should be noted that such a way of calculating the fraction of

the NEPT that reaches the sediment has associated a large uncertainty (±δ).

During the strong and weak upwelling periods, a large fraction of NEPT

sank to the bottom (75 ± 35% on 15-18 Jul and 97 ± 27 % on 22-26 Jul).

The relative importance of accumulation versus export of DON and

PON has also been assessed in each period. During the strong summer

upwelling (15-18 Jul) the decrease in DON was due to export (71%) and

oxidation (29%), whereas 63% of the PON exported to the shelf surface

waters was produced in situ and the remaining 37% came from materials

previously accumulated in the ría. During the summer relaxation (18-22

Jul) progressive accumulation of DON and PON occurred: 68% of the

DON and 61% of the PON accumulated in the study volume came from in

situ consumption of NT and the remaining 32% and 39% of DON and PON

came from allocthonous materials. Finally, during the weak summer

upwelling (22-26 Jul) 69% of the DON exported was produced in situ and

the remaining 31% of DON exported to the shelf came from the materials

previously accumulated in the ría. In contrast, 78% of the PON

accumulated previously was exported and the remaining 22% was

oxidized inside the system.

September 2002

During the autumn downwelling the ría acted as a nutrient source

(o – i = 119 ± 9 mg N m-2 d-1) at the expenses of the nutrients previously

accumulated in the system, whereas NEPT was balanced (-18 ± 12 mg N m-

2 d-1). However, during the subsequent relaxation period, the ría showed a

heterotrophic metabolism (NEPT > 0) and, 28% of NT regenerated was

exported to the shelf (o – i = 19 ± 1) and 72% was accumulated into the ría.

Page 252: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

234

Peri

od

Inte

rval

o-

i (N

T)

V*∆

(NT)/∆t

) N

EPT

(o –

i)D

(D

ON

) V

*∆(D

ON

)/∆t

) N

EPD

(D

ON

) (o

– i)

P

(PO

N)

V*∆

(PO

N)/∆

t)

NEP

P

(PO

N)

NEP

SED

St

rong

sum

mer

up

wel

ling

15 Ju

l-18

Jul

-289

±17

162±

21

-127

±37

61±8

-8

5±2

-24±

6 91

±6

-34±

18

57±2

4 95

±44

Sum

mer

rel

axat

ion

18 Ju

l-22

Jul

101±

5 -1

36±1

-3

5±5

-13±

4 41

±11

28±1

5 -2

1±3

35±2

2 14

±25

-6±3

0

Wea

k su

mm

er u

pwel

ling

22 Ju

l-26

Jul

-280

±14

121±

20

-159

±34

21±7

-7

±8

15±1

4 36

±6

-46±

17

-10±

24

154±

44

Aut

umn

dow

nwel

ling

17 S

ep-1

9 Se

p 11

9±9

-137

±4

-18±

12

-9±1

3 -1

55±5

-1

63±8

-9

4±8

110±

16

16±2

5 16

5±29

A

utum

n tr

ansi

tion

19 S

ep-2

3 Se

p 19

±1

47±1

4 66

±14

1±1

49±1

0 50

±11

-15±

2 32

±22

16±2

4 -1

32±3

0 A

utum

n up

wel

ling

23 S

ep-2

6 Se

p -9

6±8

70±1

5 -2

6±23

-2

9±9

68±9

39

±18

39±6

-6

8±14

-2

9±21

15

±36

Tabl

e 3

Net

bud

get o

f out

puts

min

us in

puts

(o –

i), a

ccum

ulat

ion

(V·∆

N/∆

t), a

nd n

et e

cosy

stem

pro

duct

ion

(NEP

) and

sed

imen

tatio

n (S

ED) o

f NT,

DO

N a

nd P

ON

in Ju

ly a

nd S

epte

mbe

r 200

2.

Page 253: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

235

During the weak autumn upwelling, the ría acted as a nutrient trap, -96 ±

8 mg N m-2 d-1, one third lower than during the weak summer upwelling.

Only 27% of NT was consumed to produce organic nitrogen and the

remaining 73% accumulated into the system. As a result, the system

changed from heterotrophic to balanced conditions (NEPT~ 0). During

September, the % of NEPT due to NO3- and NH4* was 45 and 55%

respectively, in average.

On the other hand, the relevance of the accumulation terms was also

reflected when the production of the dissolved and suspended fraction of

organic matter (NEPD and NEPP) was assessed. o – i showed that the

system act normally as a source of organic matter under upwelling

conditions (o – i > 0) and as a trap under downwelling or relaxation

conditions (o – i < 0). However, it is necessary to compare with the

accumulation of DON and PON to establish the metabolic condition of the

system independently of the hydrodynamic conditions (Table 3). During

the autumn downwelling (17-19 Sept), there was a net consumption of

DON; only 5% was imported and 95% came from the material previously

accumulated into the system. However the particulate fraction was

integrally accumulated (15% of PON was produced in situ and 85% of

PON was imported from the shelf). During the autumn relaxation (19-23

Sept) nearly 100% of the DON and 50% of the PON accumulated in the

study volume was produced in situ. Finally during the weak autumn

upwelling (23-26 Sept), DON accumulated in the study volume: 42%

imported from the shelf and 58% produced in situ). In contrast, the PON

losses were due to PON consumption (42%) and export (58%).

During the autumn downwelling, most of the NEPSED calculated with

eq. 10 came from the losses of DON previously accumulated in the

Page 254: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

236

system, since NEPT was nearly zero. Most of the DON was assimilated by

microorganisms to be transformed into PON; 10% of PON remained in

suspension (1-200 µm particle size) and 90% sedimented (>200 µm particle

size). The autumn transition was the only period when NEPSED was <0,

which means that resuspension of PON and/or diffusion of NT from the

sediments was larger than POM deposition over the bottom.

4. Discussion

4.1. Role of vertical convection and diffusion in the fertilisation of

the photic layer

The impact of shelf winds on the horizontal circulation of estuaries and

coastal inlets has been discussed thoroughly in the literature (Carter et al.

1979, Walters 1982, Wong and Garvine 1984, Wong and Moses-Hall 1998,

Janzen and Wong 2002). In the case of the Galician rías, the issue has also

been considered (Rosón et al. 1997, Álvarez-Salgado et al. 2000, Gilcoto et

al. 2001, Pardo et al. 2001, Souto et al. 2003, Piedracoba et al. 2005a). The

vertical circulation is also relevant because the fluxes resulting from the

combination of convective/diffusive velocities and the gradient of

nutrient concentrations are responsible for the fertilisation of the photic

layer. Despite that, few studies have dealt with the vertical fluxes, mainly

because vertical currents are difficult to obtain directly from

currentmeters compared with horizontal currents. In fact, most of the

studies discussing on vertical fluxes were carried out in the Galician rías

(Álvarez-Salgado et al. 1996, Rosón et al. 1999, Pérez et al. 2000b).

Water flows and nutrient fluxes in the rías were originally calculated

with a 2D steady-state box model using salinity as a tracer (Prego and

Fraga 1992). This model revealed inappropriate to reproduce the transient

upwelling events occurring in the rías. Box models improved when a

Page 255: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

237

transient and salt-heat weighted version was applied to the rías of Arousa

(Rosón et al. 1997), Vigo (Álvarez-Salgado et al. 2000) and Pontevedra

(Pardo et al. 2001). Finally, a 3D version has been recently developed by

Gilcoto et al. (2006), after empirical and numerical confirmation of a 3D

residual circulation pattern in the outer part of the Ría de Vigo (Souto et

al. 2003). Since the turbulent diffusion velocities obtained by box

modelling are representative for comparable spatial and temporal scales

than the turbulent mixing velocities of this study, direct comparisons can

be made.

The mixing velocity (KZ/∆z) was nearly twice the vertical convective

velocity (VZ) during strong summer upwelling. The weakness of coastal

winds together with thermal stratification (<N> increased by 0.04 min-1

between 18 and 22 July) lead to nearly no vertical displacements during

the subsequent summer relaxation. Finally, vertical advection became

more relevant with the re-establishment of upwelling conditions, although

previous stratification reduced the mixing between layers: (KZ/∆z)/VZ ~

0.5.

The velocities calculated in this study are coherent with the results

obtained by Míguez et al. (2001a), who applied a 2D transient box model

to the middle Ría de Vigo under similar hydrodynamic conditions: an

intense upwelling episode after a wind calm in March 1994. However, the

(KZ/∆z)/Vz ratios obtained by Míguez et al. (2001a) ranged from 0.5 to 0.7

and values of KZ/∆z > Vz were not recorded ever. The 3D transient box

model of Gilcoto et al. (2006) also produced an average ratio (KZ/∆z)/Vz

of ~ 0.5 for the upwelling period in the middle Ría de Vigo.

In September, the magnitude of the downward convective velocity

during the initial downwelling was comparable to the mixing velocity,

Page 256: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

238

which became maximum on 19 September due to the progressive decrease

of stratification (<N> decreased by -0.05 min-1 between 17 and 19

September). (KZ/∆z)/Vz was comparable with the ratio calculated by

Míguez et al. (2001a) under downwelling conditions in October 1994: Vz

ranged from -3.7 and -4.7 m d-1 and (KZ/∆z) increased from 0.3 to 3.6 m d-1

in agreement with the progressive decrease of water column stability.

Gilcoto et al. (2006) obtained a ratio > 0.7 during a downwelling period in

December.

Subsequently, stratification (<N> increased by 0.01 min-1 from 19 to 23

September) reduced the mixing velocity during the autumn transition and

at the beginning of the autumn upwelling, although the decrease was not

as intense as in July. In fact, excluding 17 September, mixing velocities

were always larger than vertical convective velocities.

Average (KZ/∆z)/Vz ratios obtained by Álvarez-Salgado et al. (2000)

with a 2D transient box model for the whole Ría de Vigo were coherent

with the ratios obtained in this work. The multiparameter inverse method

applied to the Ría de Vigo during September 1990 and 1991 by Álvarez-

Salgado and Gilcoto (2004) also reproduced similar average (KZ/∆z) and

Vz in both periods: < 1.5 m d-1, i.e. low compared with the vertical

velocities calculated in this study.

Our vertical velocities were also comparable with the values obtained

in the middle Ría de Arousa for the upwelling and downwelling season of

1989 (Rosón et al. 1999). (KZ/∆z)/Vz was 0.8 for the upwelling season,

indicating that advection was the prevalent mechanism whereas in the Ría

de Vigo mixing was sometimes up to twice the advection. During an

autumn downwelling period in October 1989, the absolute value of

(KZ/∆z)/Vz was ~0.3 in the middle Ría de Arousa. On the contrary,

Page 257: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

239

(KZ/∆z) was nearly equal to Vz in the Ría de Vigo under the same

conditions. These observations suggest that turbulent mixing is more

relevant in Vigo than in Arousa. The reason behind this difference is the ~

60% longer flushing time of the Ría de Arousa, as derived from the

relationship between the volume (Arousa, 4.34 Km3; Vigo, 3.12 Km3) and

the wideness of the open boundary (Arousa, 6.7 km; Vigo, 7.7 km) in both

rías, (4.34 x 7.7)/(3.12 x 6.7). A longer flushing time together with a 30%

larger free surface area (230 versus 176 km2) favours stratification in the

Ría de Arousa as compared with the Ría de Vigo.

Vertical nutrient fluxes depend not only on the magnitude of (KZ/∆z)

and VZ, but on the vertical profiles of nutrient concentrations too. In July,

mixing controlled the vertical transport of NT; it was up to 6 times larger

by mixing than by advection during the strong summer upwelling (15-18

Jul). During the upwelling relaxation (18-22 Jul), vertical mixing fertilises

the photic layer in spite of the downward advection of NT. In contrast, the

transport of inorganic nitrogen by mixing was 1/3 than by advection over

the upwelling season in the Ría de Arousa (Álvarez-Salgado et al. 1996a).

Although mixing fluxes reduced considerably during the autumn

downwelling (17-19 Sept) and autumn transition (19-23 Sept) compared

with July, they contributed to fertilise the photic layer despite the

dominant downwelling conditions. Vertical advection and mixing

induced by the prevailing external forces (shelf winds, continental runoff)

conditions the stability of the water column. An elevated stability can

make vertical mixing negligible.

Anyway, the vertical transport of NT, i.e. NT· VZ+ Kz · (∆NT /∆z),

depended more on the nutrient profile (60 ± 20%, p < 0.08) than on the

vertical velocity (40 ± 20%, p < 0.03) with r2= 0.85, n= 6 and p< 0.03 for the

Page 258: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

240

multiple linear correlation of NT· VZ+ Kz · (∆NT /∆z) with the average NT

concentration ( TN ) and the total (advective + diffusive) vertical velocity

(VZ+ Kz /∆z).

4.2. A geochemical approach to the origin and fate of biogenic

materials

The Ría de Vigo acts as a transient biogeochemical reactor driven by a

succession of upwelling/downwelling episodes separated by short

periods of wind calm. Therefore, the metabolic condition of this marine

ecosystem can be analyzed with a transient geochemical box model. The

transient regime of the ría increases its efficiency as a nutrient trap and

allows the accumulation of organic matter recently synthesized within the

system. The efficiency as a nutrient trap is calculated from the ratio (o –

i)/i. It was higher under weak (66±3%) and strong summer upwelling

(61±3%) than under weak autumn upwelling (56±3%). Similar efficiencies

(66%) were obtained during the upwelling season in the adjacent Ría de

Arousa (Álvarez-Salgado et al. 1996b). The trapped nutrient salts can

accumulate or be consumed: 56%, 43% and 73% of the net entry of

nutrients accumulated in the system during strong and weak summer

upwelling and weak autumn upwelling, respectively. In this situation, net

uptake by phytoplankton was delayed due to the low biomass in newly

upwelled waters and the lag time for adaptation to the light conditions.

Nutrients accumulated during upwelling supports the productivity of the

ría during the subsequent relaxation, when no entry of new nutrients from

the shelf occurs. In the Ría de Arousa, most of the nutrients trapped

during the upwelling season are transformed into organic matter (87%).

Hydrodynamic accumulation gains importance in the outer ría: up to 40%

(Álvarez-Salgado et al. 1996b).

Page 259: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

241

During downwelling/relaxation events, the ría acts as a nutrient

source. During the summer relaxation, there was a net loss of the

previously accumulated N–nutrients (74% was exported to the shelf and

the remaining 26% was consumed by phytoplankton). During the autumn

downwelling, there was also a net N–nutrients loss (87% exported, 13%

consumed) . In contrast, during the autumn relaxation, 28% of the

nutrients regenerated into the system were exported and the remaining

72% were accumulated into the system.

Under strong upwelling, phytoplankton in the photic layer is exported

to the shelf. It is replaced by the slow-growing cells in the upwelled

subsurface water, not adapted to the light conditions of the photic layer

(Zimmerman et al. 1987), and most of the upwelled N-nutrients

accumulate in the photic layer rather than being consumed. For this

reason, nutrient uptake by phytoplankton and, therefore, Chl a and PON

accumulation, use to occur during the spin down phase of upwelling

events, when nutrients are abundant, horizontal transports are reduced

and phytoplankton cells are acclimatized (Zimmerman et al. 1987,

Álvarez-Salgado et al. 1996b, Pérez et al. 2000b). The sampling frequency

(twice a week) allowed us to test that DON and PON accumulates during

upwelling relaxation, in agreement with the observations of DOM

accumulation during the decline of phytoplankton blooms observed in

other marine systems (Kirchman et al. 1994, Norrman et al. 1995, Chen et

al. 1996).

Average NEPT calculated with transient box models is suitable to

quantify the net trophic balance of a transient ecosystem (Smith and

Hollibaugh 1997, Pérez et al. 2000b, Álvarez-Salgado et al. 2001).

According to our calculations, NEPT was negative during July, showing

Page 260: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

242

an autotrophic status at the ecosystem level. In September, only the

autumn transition was strictly heterotrophic (NEPT > 0), whereas the rest

of the study period was nearly balanced (NEPT ~ 0). The geochemical

budget for July 2002 showed a net autotrophic metabolism coinciding

with the results of Álvarez-Salgado et al. (2001) during summer 1997. The

differences between the two periods are related to the relevance of

accumulation versus export of organic matter during July 2002, whereas

in July 1997 there was a close coupling between hydrodynamic and

biogeochemistry resulting that the accumulation term was nearly zero.

Autumn 1997 (Álvarez-Salgado et al. 2001) was comparable with

September 2002. In both periods, the NEPT depended mostly on the

nutrient stock previously accumulated in the ría.

The periods from 15 to 26 July and from 17 to 26 September can be

considering representative for the typical upwelling-relaxation and

downwelling-relaxation events off NW Spain, respectively. Average NEP

during July was 107 mg N m-2 d-1 (= TNEP− ), very close to the estimation

of Gilcoto et al. (2001) for the same area (140 mg N m-2 d-1, assuming C/N

= 6.7) during an upwelling relaxation in September 1990. About 25% of

this material was exported to the shelf. In September, only the weak

autumn upwelling period was productive ( TNEP = -26 mg N m-2 d-1), and

the fraction exported to the shelf was about 40%. These results are

coherent with Gago et al. (2003) who estimated that 32% of the organic

carbon was exported to the shelf. The estimations are also comparable

with the values given by Waldron et al. (1998) for the Benguela upwelling

system, where 34% of the new production was exported offshore. In the

Ría de Arousa, 80% of NEP was exported to the shelf during the

upwelling season of 1989 (Rosón et al. 1999), a percentage that reduces to

Page 261: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Geochemical approach to the physical–biological coupling in the Ría de Vigo

243

50% when only the middle ría is considered (Álvarez-Salgado et al.

1996a). Regarding the POM:DOM ratio of NEP, the estimations of Gago et

al. (2003) and our own results yields a PON:DON ratio of 60:40. On the

contrary, the massive culture of mussels on hanging ropes in the Ría de

Arousa causes an inverse PON:DON ratio of 30:70 (Álvarez-Salgado et al.

1996a). The results of our work indicates that the PON:DON ratio of the

exported material was independent of the upwelling intensity. However,

Gago et al. (2003) found that this ratio varied according to the upwelling

intensity during July 1997. During the weak autumn upwelling, only PON

was exported to the shelf. In contrast, DON was imported from the shelf

and accumulated into the system.

The NEP of DON (NEPD) is coupled with the balance of outputs minus

inputs of DON (o-i)D only under weak summer upwelling. The ría acts as

a DON source during moderate upwelling, as reported by Gago et al.

(2003). The importance of the accumulation term during the rest of July

and September leads to a decoupling of the production and export of

DON and PON. Our results show that most of the material exported to the

shelf comes from losses of the material previously accumulated rather

than recently produced.

Page 262: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 5

244

Summary

Lateral differences in the distributions of N-nutrients, PON and DON

were observed in the middle Ría de Vigo, depending on the season and

the specific hydrographic and dynamic conditions. The lateral variability

of both the N species concentrations and the longitudinal component of

the residual current led to significant lateral differences in the N-nutrients,

PON and DON fluxes and, therefore, in the corresponding i – o balances.

The hydrography (stratification vs homogenisation) and dynamics

(upwelling vs downwelling) condition the relative importance of vertical

convection and turbulent mixing in the fertilisation of the photic layer.

Mixing became the most important fertilisation mechanism under strong

upwelling conditions and compensated the downward vertical convection

during downwelling episodes, ensuring the fertilisation of the photic

layer.

The need to assess the relative importance of the production and

accumulation of organic matter in a transient marine ecosystem makes the

geochemical approach followed in this paper the most suitable to

characterise the metabolic state of the Ría de Vigo at the ecosystem level.

Page 263: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6:

Physical-biological coupling in the

coastal upwelling system of the Ría de Vigo:

An in vitro approach

Page 264: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 265: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

247

Chapter 6, the research work presented in this chapter is also a

contribution to the paper:

S. Piedracoba, M. Nieto-Cid, I.G. Teixeira, J.L. Garrido, X.A.

Álvarez-Salgado, G. Rosón, C.G. Castro, F.F. Pérez. 2005. Physical-

biological coupling in the coastal upwelling system of the Ría de Vigo

(NW Spain). I: An in vitro approach. Mar. Ecol. Prog. Ser. (submitted).

Abstract: The metabolism of the pelagic ecosystem of the Ría de Vigo

(NW Spain) was studied during July and September 2002. Measurements

of oxygen production (Pg) and respiration (R) were performed in a single

site (water depth, 40m) twice a week at five depths, from surface to

bottom. In July, the net community production (NCP = Pg – R) reduced in

parallel with the decrease of Pg, from 0.73 g N m-2 d-1 under strong

upwelling to 0.14 g N m-2 d-1 under weak upwelling conditions. In

September, the NCP increased from -0.10 g N m-2 d-1 under downwelling

to 0.0 g N m-2 d-1 under weak upwelling conditions. In addition,

microzooplankton grazing and sedimentation rates were measured for

first time in the Ría de Vigo. The high grazing rates observed questions

the efficiency of the ría to transfer organic matter directly from

phytoplankton to the metazoans. In fact, organic matter should be

imported from the shelf to support the metabolic requirements of the

microheterotrophs in September. Comparison of the metabolic state of the

Ría de Vigo derived from these in vitro measurements (Pg, R, grazing and

sedimentation) and the in situ geochemical budget of the companion

paper by Piedracoba et al. (Chapter 5) produce contrasting results: they

agree only in 3 of the 6 study cases. It is concluded that both methods are

Page 266: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

248

complementary and their simultaneous application allows to obtain a

better knowledge of the system functioning.

Page 267: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

249

Resumen: Se ha estudiado el metabolismo del ecosistema pelágico de

la Ría de Vigo (NO de España) durante los períodos de julio y septiembre

de 2002. Para ello se midieron dos veces por semana las tasas de

producción (Pg) y respiración (R) en cinco niveles entre superficie y fondo

en una estación de 40 m de profundidad. En julio, la producción neta de

comunidad (NCP = Pg – R) disminuyó con Pg, desde 0.73 g N m-2 d-1 bajo

condiciones de afloramiento intenso a 0.14 g N m-2 d-1 bajo condiciones de

afloramiento moderado. En septiembre, la NCP aumentó desde -0.10 g N

m-2 d-1 bajo condiciones de hundimiento a 0.0 g N m-2 d-1 bajo condiciones

de afloramiento moderado. Además, se han medido por primera vez en la

Ría de Vigo, tasas de herbivoría y de sedimentación. Las altas tasas de

herbivoría cuestionan la eficiencia de la ría para transferir materia

orgánica directamente desde el fitoplancton hasta los metazoos. De hecho,

durante el mes de septiembre la ría necesitó importar materia orgánica

desde la plataforma para soportar los requerimientos metabólicos de los

microheterótrofos. Se ha comparado el estado metabólico de la Ría de

Vigo derivado de las medidas in vitro (Pg, R, herbivoría y sedimentación)

con el obtenido a través de un balance geoquímico (parte I de los dos

manuscritos). Los resultados son coincidentes en 3 de los 6 casos de

estudio. Se concluye que ambos métodos son complementarios y su

aplicación simultánea permite conocer mejor cómo funciona el sistema.

Page 268: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

250

Page 269: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

251

1. Introduction

This work is the companion paper of Piedracoba et al. (Chapter 5), in

which the new production (NP) of the central segment of the transient

coastal upwelling system of the Ría de Vigo was evaluated at the short

time scale (< 4 days) under dominant upwelling (in July) and

downwelling (in September) conditions using a mass balance box model.

Although NP rates estimated with the mass balance approach are

ready for direct interpretation at the ecosystem level (Smith and

Hollibaugh 1997), this method does not provide any insight on the specific

biogeochemical processes (production, respiration, sedimentation,

grazing, etc…) that lead to a given NP under a particular hydrographic

and dynamic scenario. Therefore, part II will be devoted to the evaluation

of the short time scale variability of different biogeochemical rates

estimated with sediment traps and in vitro incubations (net community

production and dark respiration by the oxygen dark/light incubation

method and microzooplankton herbivory by the dilution technique).

Fluxes derived from the individual in vitro determinations, some of them

evaluated for the first time in the Ría de Vigo, are then summed to derive

the metabolic status of the system. As pointed out by Smith and

Hollibaugh (1993; 1997), there are potential problems with using this

approach to assess net metabolism: i) analytical artefacts associated with

isolating a portion of the community in containers and performing such

incubation experiments; ii) some of the components can be left out of the

summation; iii) the anaerobic benthic metabolism is only considered by

CO2 based measurements, but it is missed by O2 based measurements; or

iv) method-dependent biases between the measurements of each

component.

Page 270: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

252

Finally, the results of the mass balance approach of Piedracoba et al.

(Chapter 5) and the in vitro approach presented in this work, applied

simultaneously for the first time in the coastal upwelling system of the Ría

de Vigo, will be compared and discussed in terms of their respective

strengths and weaknesses.

2. Material and methods

Stn 00 (Figure 1), in the middle segment of the coastal upwelling

system of the Ría de Vigo was visited about 1 hour before sunrise on 15,

18, 22 and 26 July and 17, 19, 23 and 26 September 2002 (see Piedracoba et

al. Chapter 5 for details). Water samples were taken with a rosette sampler

equipped with twelve 10–litre PVC Niskin bottles with stainless–steel

internal springs. Salinity and temperature were recorded with a SBE 9/11

conductivity–temperature–depth probe attached to the rosette sampler.

Conductivity measurements were converted into practical salinity scale

values with the equation of UNESCO (1985). Water samples for the

analyses of the chemical variables and the production/respiration rates

were collected from five depths: surface (50% Photosynthetic Available

Radiation, PAR; 2.5±0.3 m), 25% PAR (7±1), 1% PAR (14.2±2.2 m),

26.2±1.2m and bottom (40.3±0.5 m). Microzooplankton grazing

experiments were performed only at the surface (50% PAR).

Chemical variables. Salinity (S, accuracy ± 0.003), dissolved oxygen

(O2, ± 0.5 µmol kg–1), N-nutrient salts (NH4+, ± 0.05 µM; NO2–, ± 0.02 µM;

and NO3–, ± 0.1 µM), dissolved organic nitrogen (DON, ± 0.2 µmol N L-1),

particulate organic nitrogen (PON, ± 0.1 µmol N L-1) particulate

carbohydrates (PCho, ± 0.1 µmol C L-1), chlorophyll a (Chl, ± 0.05 mg m-3)

and pigments have been measured as described elsewhere (Álvarez-

Salgado et al. 2005, Gago et al. 2005).

Page 271: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

253

Chemical composition of particulate organic matter. Assuming that

the changes in the C/N/P ratio of particulate organic matter (POM) are

due to variations in the proportions of the major groups of biomolecules

rather than in the chemical formula of each group, it is possible to

estimate their proportions. Fraga et al. (1998) reviewed the average

composition of marine phytoplankton carbohydrates (PCho, C6H10O5),

lipids (Lip, C53H89O6), proteins (Prt, C139H217O45N39S), phosphorus

compounds (Pho, C45H76O31N12P5) and pigments (Chl, C46H52O5N4Mg).

From the chemical formula of these biomolecules, the following set of five

linear equations can be written:

Chl46Pho45Lip53PCho6tPr139POC ×+×+×+×+×= (1)

Chl52Pho76Lip89PCho10tPr217H ×+×+×+×+×= (2)

Chl5Pho31Lip6PCho5tPr45O ×+×+×+×+×= (3)

Chl4Pho12tPr39PON ×+×+×= (4)

Pho5POP ×= (5)

Since POC, PON, POP, PCho and Chl were measured, the system can

be solved to obtain the average chemical formula, and the proportions of

the different biomolecules for each particular sample. The five unknowns

are H, O, Prt, Lip and Pho.

Metabolic balance of the water column. Daily photosynthetic (Pg) and

respiration (R) rates of the plankton community were estimated by the

oxygen light–dark bottle method (Strickland and Parsons 1972). Samples

collected before sunrise in 10 L Niskin bottles were transferred to black

polyethylene carboys. The carboys were gently shaken before sampling to

prevent sedimentation of the particulate material. Series of eleven 110 mL

Winkler bottles composed of triplicate initial, and quadruplicate light and

Page 272: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

254

dark subsamples were filled. Each series of bottles were incubated for 24

hours (starting within 1 hour of the sun rise) at the original light and

temperature conditions in an incubator placed in the terrace of the base

laboratory. In situ conditions were reproduced in the incubators by

mixing the appropriate proportions of the cold (10°C) and warm (30°C)

water provided by the cultivation facility of the base laboratory and using

increasing numbers of slides of a 1 mm mesh, respectively. Dissolved

oxygen was determined by Winkler potentiometric end–point titration.

Microzooplankton herbivory. Four experiments, on 18 and 26 July and

19 and 26 September, were carried out using the dilution technique

(Landry and Hassett 1982). Surface water was collected with a 30 L Niskin

bottle dipped twice. Water from the first dip was gravity filtered through

a 0.2 µm Gelman Supocap filter and mixed with unfiltered seawater from

the second dip in twelve 2.3 L clear polycarbonate bottles to obtain

duplicates of 10, 20, 40, 60, 80 and 100% of plankton ambient levels.

Bottles were completely filled and incubated for 24 hours at the original

light and temperature conditions in an incubator placed in the terrace of

the base laboratory. Initial and final subsamples were taken from all

experimental bottles for the determination of carbon biomass of pico-,

nano- and microplankton. For pico- and nanoplankton carbon biomass

determination, subsamples of 10mL fixed with buffered formaldehyde

(2% final concentration) and stained with DAPI, were filtered through

25mm black polycarbonate filters (0.2µm of pore size) and examined using

an epifluorescence Nikon microscope (Porter and Feig 1980). For

microplankton carbon biomass determination, subsamples preserved in

Lugol’s iodine and sedimented in composite sedimentation chambers

were observed through an inverted microscope (Utermohl 1958).

Page 273: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

255

Dimensions of the species or groups identified were measured and cell

volumes were converted to cell carbon according to Strathmann (1967) for

diatoms and dinoflagellates, Verity et al. (1992) for flagellates and Putt

and Stoecker (1989) for ciliates. Changes in carbon biomass concentration

were used to estimate the net community growth rate (k):

⎟⎟⎠

⎞⎜⎜⎝

⎛⋅=

0

t

CC

lnt1k (6)

Where t is the duration of the experiment (1 d); and C0 and Ct are the

initial and final carbon biomass concentrations, respectively.

Instantaneous growth (µ) and grazing (g) rates were calculated by the

linear regression of the daily net growth rate (κ) against the dilution factor

(D):

gD ⋅−µ=κ (7)

In cases of saturated feeding responses, µ was obtained by the

regression of the linear part of the curve and g was calculated by the

difference between the net growth rate in the undiluted sample and µ

(based on Gallegos 1989).

Sedimentation rates. A multitrap collector was deployed at the 1%

PAR, and anchored to the seabed, for about 6 h on 18 and 22 July and 19

and 26 September. The collector was described in Knauer et al. (1979)

following the JGOFS protocols (UNESCO 1994) and consisted of 4

individual multitrap baffled cylinders of 6 cm diameter and 60 cm length.

Each cylinder was placed at the middle of a PVC cross. Prior to each

deployment, the cylinders were filled with a solution of filtered seawater

to which 5 g of NaCl was added per litre of solution. This solution

prevented the exchange of material with the surrounding water. The

Page 274: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

256

9.0º 8.9º 8.8º 8.7ºLongitude (West)

42.1º

42.2º

42.3º

Latit

ude

(Nor

th)

Rande Strait San

Sim

on B

ay

C. MarCíes Islands

P.Borneira

DCM12 st.&

00

Ría de Vigo

rosette station

content of each cylinder was collected on 47 mm pore size Whatman GF/F

filters to determine POC, PON, POP and Chl.

Figure 1. Map of the Ría de Vigo (NW Spain). Stn 00, where the ADCP was moored and water samples were taken, is shown.

3. Results

The four sampling dates in July and September were grouped

according to the observed hydrodynamic conditions (Piedracoba et al.

Chapter 5). Three periods were defined in July: a) “spin down of strong

summer upwelling” (15-18 Jul), b) “summer relaxation” (18-22 Jul) and c)

“spin up of weak summer upwelling” (22-26 Jul). In September, the three

intervals were: d) “autumn downwelling” (17-19 Sep), e) autumn

transition” (19-23 Sep) and f) “autumn upwelling” (23-26 Sep).

3.1. Chemical composition of the biogenic materials

The contribution (%mol C/mol C) of proteins + phosphorous

compounds (i.e. the labile N/P compounds), carbohydrates and lipids to

POM in the photic and aphotic layers is shown in Table 1. The proportion

of N/P compounds did not change significantly with depth: it ranged

Page 275: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

257

from 60-65 % in the upper to 57-64% in the lower layer. In general, July

was characterized by an increase of lipids and a decrease of carbohydrates

from surface to bottom. During the strong and weak summer upwelling,

there was an increase of N/P compounds and lipids at the expenses of a

decrease of carbohydrates. The contrary was observed during the summer

relaxation period. In general, July was characterized by a progressive

increase of the most recalcitrant material (lipids) with the weakness of

upwelling due to preferential mineralization of the most labile

compounds in shallower levels (Torres-López et al. 2005). In September,

the labile N/P compounds ranged from 50 to 60 %, increasing from

surface to bottom as a consequence of the combination of downwelling

and fast sinking of fresh POM from the surface layer. In addition, this

nutrient-limited period was characterized by a high proportion of

carbohydrates (Table 1).

The stoichiometric ratios of conversion of dissolved oxygen production

into organic carbon (Rc), nitrogen (Rn) and phosphorus (Rp) for the upper

and lower layers were obtained from the corresponding chemical

formulas of POM (Table 1). These stoichiometric ratios will be used to

transform into an homogeneous currency (nitrogen) the biogeochemical

rates obtained in different unities.

Page 276: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

258

Date Pho Prot Lip Cho formula O2 Rc Rn Rp 15-7 U 19.14 46.36 18.85 15.64 C67H108O25N11P 96 1.4 9 96 18-7 U 16.21 50.36 23.79 9.64 C82H132O27N13P 119 1.5 9.1 119 22-7 U 13.64 38.77 15.9 31.69 C91H148O40N12P 125 1.4 10 125 26-7 U 13.55 49.69 14.57 22.19 C93H149O37N15P 132 1.4 8.8 132 17-9 U 14.13 44.59 18.91 22.36 C91H146O35N13P 128 1.4 9.6 128 19-9 U 6.46 44.85 9.24 39.46 C189H303O90N26P 254 1.4 9.7 254 23-9 U 13.7 38.53 10.83 36.94 C88H142O43N12P 119 1.4 9.9 119 26-9 U 11.75 47.52 17.1 23.64 C109H175O43N16P 153 1.4 9.4 153 15-7 L 13.18 50.65 22.58 13.59 C100H162O34N16P 145 1.5 9.3 145 18-7 L 14.72 42.93 28.63 13.72 C92H149O30N12P 131 1.4 11 131 22-7 L 15.81 41.66 22.61 19.91 C82H133O31N11P 116 1.4 10 116 26-7 L 14.2 44.94 26.91 13.95 C95H153O31N13P 135 1.4 10 135 17-9 L 14.33 42.36 25.01 18.3 C92H150O33N13P 131 1.4 10 131 19-9 L 17.86 35.87 14.1 32.16 C68H111O32N9P 93 1.4 10 93 23-9 L 8.48 48.63 10.58 32.31 C145H233O65N22P 201 1.4 9.1 201 26-9 L 14.03 39.29 9.93 36.75 C86H138O42N12P 116 1.4 9.6 116

Table 1. Percentages of P compounds (Pho), proteins (Prot), lipids (Lip) and carbohydrates (Cho) in the particulate organic material of samples collected at stn 00. Average composition of the biogenic materials in the upper (U; photic) and lower (L; aphotic) layers of the middle segment of the Ría de Vigo in July and Sptember 2002. The stoichiometric ratio of conversion of O2 into organic carbon (Rc

=∆Corg∆O2 ), organic nitrogen (Rn =

∆Norg∆O2 ) and organic phosphorous ( Rp =

∆Porg∆O2 ) are provided.

A preferential removal of nitrogen and phosphorous relative to carbon

during the initial stages of particulate organic matter decomposition

occurred. The fractionation of the stoichiometry of the mineralization of

POM, and also DOM, was showed previously by Martin et al. (1987),

Minster and Boulahdid (1987), Shaffer et al. (1999), Brea et al. (2004), in the

open sea, and recently by Álvarez-Salgado et al. (2006) in shelf waters.

The stoichiometric ratios of conversion of dissolved oxygen production

into organic carbon (Rc), nitrogen (Rn) and phosphorus (Rp) for the upper

and lower layers were obtained from the corresponding chemical

formulas of POM (Table 1). These stoichiometric ratios will be used to

Page 277: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

259

transform into a homogeneous currency (nitrogen) the biogoechemical

rates obtained in different unities.

3.2. Metabolic balance of the microbial loop in relation to

phytoplankton populations

Pg reached a surface maximum on 18 July (Figure 2c, Table 2), during

the strong summer upwelling, coinciding with the positive residual

circulation pattern (Figure 2b) promoted by shelf northerly winds (Figure

2a). This Pg maximum was accompanied by high Chl a and POC levels

(Figure2e, Figure 2 in Piedracoba et al. Chapter 5). The HPLC analysis of

phytoplankton pigments revealed the predominance of species in the

micro- and nanoplankton fraction (> 5 µm); picoplankton (< 5 µm) never

accounted for >5 % of total Chl a. The pigment composition in the micro-

and nanoplankton fraction was characterized by high levels of

fucoxanthin (ranging from 0.36 to 8.30 µg·L-1, Figure 2f). Neither acyloxy-

fucoxanthin derivatives nor chlorophyll c-galactolipid esters were

detected in more than trace amounts, but chlorophyll c3 (Figure 2g) was

relatively abundant during this period. This pigment signature suggests

the importance of chlorophyll c3 -containing diatom species. Cell counts

for samples taken on 18 July found more than 4 x 106 cells L-1 of

Leptocylindrus danicus.

During the summer relaxation period, Pg decreased (268 mmol O2 m-2

d-1) and a subsurface R maximum (222 mmol O2 m-2 d-1) at 20 m depth

(Figure 2d) was recorded. It coincided with a subsurface chlorophyll a and

fucoxanthin maximum (Figures 2e, f), suggesting that it corresponded to

the same diatom population that caused the surface maximum of Pg

during the previous strong summer upwelling. In any case, the water

column integrated R was lower than Pg and, therefore, the system was

Page 278: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

260

autotrophic. Significant amounts of chlorophyllide a (up to 1.33 mg L-1 in

surface samples on 18 July) were found during this period. The

distribution of chlorophyllide a (Figure 2h) followed the same vertical

distribution of chlorophyll a and fucoxanthin, indicating that the presence

of this degradation pigment can be also linked to the same phytoplankton

group. However, the vertical distribution of chlorophyll c3 did not match

exactly the same vertical gradient (Figure 2g). The occurrence of small

amounts of peridinin (ranging from 0.09 to 0.77 µg L-1) indicates the

presence of type 1 dinoflagellates (sensu Jeffrey and Wright 2006).

During the weak summer upwelling period, the positive residual

circulation pattern promoted the entry of nutrient-rich waters (Figure 2 in

Piedracoba et al. Chapter 5) that caused a marked Pg increase, although

lower than during the strong summer upwelling period. Higher R values

reached shallower waters in relation to the previous period. Pg and R

maxima occurred at the same level.

The photic-zone-integrated f-ratio, (Pg – R)/Pg (Quiñones and Platt

1991), decreased from the strong (87%) to the weak summer upwelling

(40%), showing that the fraction of the phytoplankton production

available to be exported either horizontally or vertically or to be

transferred to higher trophic levels tended to decrease along the study

period.

During the autumn downwelling (17-19 Sept), the system was balanced

(Pg ~ R, Figures 3c, d, Table 2), coinciding with very low Chl and nutrient

levels (Figure 3e, Figure 4 in Piedracoba et al. Chapter 5). The size

distribution of pigments revealed a higher contribution of picoplankton,

which was specially significant in subsurface samples (Figure 3e): 16.1 and

29.7 % of total chlorophyll a at 25 m depth on 17 and 19 September,

Page 279: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

261

respectively. The picoplankton pigment composition of these samples was

characterized by the occurrence of zeaxanthin and chlorophyll b,

suggesting the presence of cyanobacteria. The contribution of

picoplanktonic Chl a to total Chl a was <6% in the remaining samples.

The pigment composition of micro- and nanoplankton was

characterized by a high contents of peridinin (up to 11.28 µg L-1, Figure

3f), indicating the predominance of type 1 dinoflagellates (Jeffrey and

Wright 2006). Smaller amounts of fucoxanthin (ranging from 0.08 to 1.14

µg L-1) were found with no acyloxy-fucoxanthin derivatives, chlorophyll c3

or non-polar chlorophyll c derivatives, thus indicating the occurrence of

diatoms (Jeffrey and Wright 2006). Alloxanthin, a marker pigment for

cryptophytes, was detected also in small amounts (from 0.04 to 0.31 µg L-1,

Figure 3g). Cell counts on 19 September samples found 0.6 x 106 cells L-1 of

Cryptophyceae sp.

During the beginning of the transition period, Chl increased in the

upper layer but Pg was low compared with R. The vertical distribution of

peridinin was restricted to the photic zone (within the first 15 m of the

water column), and reached a maximum at the 50 % PAR on 19 September

(Figure 3f). Fucoxanthin, with much smaller but rather uniform amounts,

was more widely distributed both in depth and in time. As in July,

chlorophyllide a was detected, but in much smaller amounts (up to 0.28

µg L-1). Chlorophyllide a reached a maximum at the 50 % PAR on 23

September, but significant amounts were also be detected at the bottom

layer the same day (Figure 3h).

When the hydrodynamic conditions changed to upwelling, a Pg

maximum occurred (277 mmol O2 m-2 d-1), although it was nearly 1/3 of

the maximum in July. On the contrary, R was similar in both periods. Cell

Page 280: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

262

counts on 26 September found 106 cells L-1 of Skeletonema costatum,

Chaetoceros spp and Thalassiosira nana. The contribution of picoplankton to

total Chl a was 14.1 % at the surface on 26 September. Alloxanthin

appeared near the surface on 17 and 19 September, but the situation

reversed as the month goes by, with highest values detected at the lowest

depths sampled on 23 and 26 September (Figure 3g), indicating that

cryptophytes were the main contributors to the chlorophyll a measured

there (Figure3e).

P R

upper lower upper lower

15/07/2002 (mmol/(m2d) 518 0 80 85 (g N/(m2d) 0.81 0 0.1 0.13 18/07/2002 (mmol/(m2d) 697 0 75 49 (g N/(m2d) 1.07 0 0.12 0.06 22/07/2002 (mmol/(m2d) 268 0 222 89 (gN/(m2d) 0.36 0 0.34 0.12 26/07/2002 (mmol/(m2d) 365 0 95 51 (gN/(m2d) 0.58 0 0.13 0.07 17/09/2002 (mmol/(m2d) 21 0 56 78 (gN/(m2d) 0.03 0 0.09 0.1 19/09/2002 (mmol/(m2d) 119 0 61 75 (gN/(m2d) 0.17 0 0.09 0.1 23/09/2002 (mmol/(m2d) 277 0 135 89 (gN/(m2d) 0.39 0 0.19 0.14 26/09/2002 (mmol/(m2d) 84 0 60 66 (gN/(m2d) 0.13 0 0.08 0.09

Table 2. Depth average gross (Pg) primary production and respiration (R) rates in O2 and nitrogen units for the upper and lower layer of the middle segment of the Ría de Vigo during the sampling dates. Rn of the photic (upper) and aphotic (lower) layer was used to convert O2 into N units (Table 1).

Page 281: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

263

Figure 2. . Time course of shelf winds, in m s–1 (a); residual currents, in cm s–1 (b); gross primary production, in µmol O2 kg–1 d–1 (c); respiration, in µmol O2 kg–1 d–1 (d); chlorophyll a, in µg L–1 (e); fucoxanthin, in µg L–1 (f); chlorophyll c3, in µg L–1 (g); and chlorophyllide a, in µmol L–1 (h) during July 2002.

-40

-30

-20

-10

15/7

18/7

22/7

26/7

c

d

b

-40

-30

-20

-10

e

f

g

-40

-30

-20

-10

15/7

18/7

22/7

26/7

-40

-30

-20

-10

-40

-30

-20

-10

h

-10

-20

-30

-40

-30

-20

-10

-12-606

12

16/7 18/7 20/7 22/7 24/7 26/7

a

Page 282: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

264

Figure 3. Time course of shelf winds, in m s–1 (a); residual currents, in cm s–1 (b); gross primary production, in µmol O2 kg–1 d–1 (c); respiration, in µmol O2 kg–1 d–1 (d); chlorophyll a, in µg L–1 (e); peridinin, in µg L–1 (f); alloxanthin, in µmol L–1 (g); and chlorophyllide a, in µmol L–1 (h) during September 2002.

17/9

19/9

23/9

26/9

b

c

d

-40

-30

-20

-10

-40

-30

-20

-10

h17/9

19/9

23/9

26/9

-40

-30

-20

-10

-40

-30

-20

-10

-40

-30

-20

-10

e

f

g

-10

-20

-30

-40

-30

-20

-10

-12-606

12

17/9 19/9 21/9 23/9 25/9

a

Page 283: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

265

The photic-zone-integrated f-ratio ranged between -59% during the

autumn downwelling to 40% during the weak autumn upwelling. The

negative sign shows that the system was heterotrophic (R > P).

3.3. Microzooplankton grazing

The four dilution experiments produced contrasting results: µ = 1.32 ±

0.24 d-1 (average ± standard error) and g = 1.04 ± 0.30 d-1 on 18 July.

Similar values were obtained on 26 July: µ = 1.43 ± 0.10 d-1 and g = 0.85 ±

0.20 d-1. On the contrary, µ and g were considerably lower in September: µ

= 0.36 ± 0.02 d-1 and g = 0.31 ± 0.10 d-1 on 19 September and µ = 0.32 ± 0.2

d-1 and g = 0.13 ± 0.20 d-1 on 26 September.

A constant with depth g was assumed on the basis that

microheterotrophs and phytoplankton species composition do not change

significantly through the photic layer (Nogueira et al. 2000). The photic-

layer-averaged growth rate, <µ>, was calculated from the µ of the 50%

PAR and the proportion between the 50% PAR ( %50Pg , mmol O2 m-3 d-1)

and the photic layer integrated O2 primary production (<Pg>, mmol O2 m-

2 d-1):

µ⋅⋅

=µzPg

Pg

%50 (8)

where z ( in m) is the 1% PAR depth. The resultant <µ> were 0.82 ± 0.06

d–1 on 18 July, 0.65 ± 0.03 d–1 on 26 July, 0.20 ± 0.01 d–1 on 19 September

and 0.24 ± 0.08 d–1 on 26 September.

Some more assumptions are needed to calculate <µ> and g the dates

when grazing experiments were not carried out. The g/µ ratio at the 50%

PAR measured on 18 July was applied to 15 and 22 July. In September, the

g/µ ratio measured on 19 and 26 September was applied to 17 and 23

September, respectively. These assignments were based on the similarity

Page 284: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

266

of the hydrographic and dynamic conditions among sampling dates. Once

the g/µ ratios have been set, µ for 15 July, 22 July, 17 September and 23

September were calculated with the formula of Calbet and Landry (2004)

by finding the value of µ that produce the measured Pg50%:

g1eChl

ChlOPg

g

02

%50 −µ−

⋅⋅µ⋅⎟⎠⎞

⎜⎝⎛∆∆

=−µ

(9)

Where Chl0 is the initial Chl a concentration and ⎟⎠⎞

⎜⎝⎛∆∆ChlO 2 is the

stoichiometric ratio of conversion of Chl a into O2 changes, obtained from

the average composition of POM in the photic layer each sampling date

(Table 1). To test the confidence of eq. 13 for the calculation of µ and g, it

was solved for the dates when grazing experiments were conducted; this

test yields that the difference between the measured and calculated Pg50%

was less than ±20%.

The values of µ and g obtained were 1.89 ± 0.19 and 1.58 ± 0.39 d-1 on

15 July; 0.40 ± 0.04 and d-1 and 0.31 ± 0.08 d-1 on 22 July; 0.17 ± 0.02 d-1 and

0.14 ± 0.02 d-1 on 17 September; and 0.41 ± 0.04 d-1 and 0.16 ± 0.06 d-1 on

23 September. Using eq. 12, it resulted that <µ> was 1.13 ± 0.07 d-1 on 15

July, 0.31 ± 0.03 d-1 on 22 July, 0.09 ± 0.03 d-1 on 17 September and 0.30 ±

0.02 d-1 on 23 September. Finally, the photic layer integrated grazing rate

(in g N m-2 d-1), G, was calculated as:

RnzPgg

G ⋅⋅⋅µ

= (10)

Table 3 summarises the direct and indirect estimates of

microzooplankton grazing parameters. The percentage of Pg uptaken by

microheterotrophs ranged between 64% and 81% during July and,

decreased from 115% to 44% during September.

Page 285: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

267

Date <Chl0> zm Chl 0 µ g Pg 50% <µ> ∆C/∆Chla ><><

PgChlag*

15/07/2002 8.88 13 10.7 1.9 1.6 23.45 1.1 2.67 80% 18/07/2002 9.82 14 18.9 1.3 1 24.88 0.8 2.9 51% 22/07/2002 6.14 21 3.06 0.4 0.3 1.26 0.3 11.14 140% 26/07/2002 4.24 17 5.56 1.4 0.9 9.22 0.7 4.46 67% 17/09/2002 2.3 11 2.22 0.2 0.1 0.37 0.1 8.61 164% 19/09/2002 4.4 12 5.71 0.4 0.3 2.4 0.2 6.76 108% 23/09/2002 6.8 15 8.06 0.4 0.2 3.73 0.3 5.66 44% 26/09/2002 1.79 12 2 0.3 0.1 0.67 0.3 11.53 45%

Table 3. Summary of the grazing incubation experiments. <Chl0> is the initial Chl a in the photic layer (in µg L–1); Chl0 is the initial Chl a in the 50%PAR (in µg L–1); zm is the depth of the 1% PAR (in m); µ is the phytoplankton gross growth rate in the 50% PAR (in d–1) ; g is the microzooplankton grazing rate in the 50% PAR (in d–1); <µ> is the average µ in the photic layer; PP50% is the O2 primary

production rate in the 50% PAR (in mmol O2 m-3 d-1); ⎟⎠⎞

⎜⎝⎛∆∆ChlC is the stoichiometric ratio of conversion

of Chl a into C changes obtained from the average composition of POM in the upper layer each sampling date (Table 1).

3.4. Vertical fluxes of biogenic organic materials

Sedimentation rates were 0.34 ± 0.03 and 0.14 ± 0.04 g N m-2 d-1 on 18

and 22 July and 0.15 ± 0.03 and 0.17 ± 0.04 g N m-2 d-1 on 19 and 26

September. Dividing the sedimentation rates by the average PON

concentration of the photic layer, sedimentation velocities (ϖ ) of 5.7 ± 0.7

m d-1, 2.6± 0.9 m d-1, 2.4 ± 0.5 m d-1 and 3.6 ± 0.9 m d-1 were calculated.

Considering the depth where the trap was deployed (the 1% PAR),

sedimentation times of 2 d on 18 July, 8 d on 22 July, 8 d on 19 September

and 5 d on 26 September were computed. Since the sampling frequency

was, in general, lower than the sedimentation time, time-average

sedimentation velocities ϖ of 5.8 ± 0.7 m d-1 and 2.4 ± 0.8 m d-1 were

estimated for 15-18 and 18-22 July, and 2.7 ± 0.6 m d-1, 2.3 ± 0.7 m d-1 and

2.6 ± 0.7 m d-1 for 17-19, 19-23, 23-26 September, which considers the days

when sediment traps were not deployed. Vertical convective and diffusive

Page 286: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

268

PON and DON’ fluxes were also calculated in each period (Table 4).

DON’ was obtained by subtracting the average DON in oceanic ENACW

(3.6 ± 0.4 µmol L–1; Álvarez-Salgado et al. 2001, Nieto-Cid et al. 2005) to

the measured DON concentration.

In July, a decrease of the downward vertical flux of organic nitrogen

(DON’ + PON) was observed in parallel to the weakening of upwelling,

mainly due to a net reduction of turbulent mixing (Table 4) and

sedimentation rate. Sedimentation was the dominant vertical transport

process: it was responsible for 70 to 90% of the vertical flux. PON was the

prevalent form transported by vertical convection (80%), while the

percentage of DON’ and PON transported by turbulent mixing was 50:50,

except during the strong summer upwelling, when it changed to 30:70

(DON’: PON).

In September, all the material previously accumulated in the photic

layer was transported downwards during the autumn downwelling, with

vertical convection dominating over turbulent mixing. In general, the

DON’ transport was larger than in July, probably because the survey was

executed after an intense upwelling and, therefore, there was DON

accumulation in the photic layer. In fact, during the transition from

autumn downwelling to upwelling, turbulent mixing of DON’ was the

dominant vertical flux. Sedimentation rates increased with the

reestablishment of upwelling (70% of the PON was transported by

sedimentation and 30% by turbulent mixing).

Page 287: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

269

Period Interval DON’*Vz DON’*Kz/∆z PON*Vz PON *Kz/∆z

Strong summer upwelling 15-18 Jul

-17±4 141±3 -75±13 121±6

Summer relaxation 18-22 Jul

3±0 17±0 21±5 14±1

Weak summer upwelling 22-26 Jul

-19±6 13±1 -71±18 30±3

Autumn downwelling 17-19 Sep

101±2 71±3 142±3 132±3

Autumn transition 19-23 Sep

19±3 136±3 40±6 -1±4

Autumn upwelling 23-26 Sep

-68±6 71±2 -88±10 -138±2

Table 4. Vertical convective and diffusive mixing PON and DON’ fluxes (<0, upward; >0, downward) in mg N m-2 d-1.

4. Discussion

Nitrogen (in g N m-2 d-1) and carbon (in g C m-2 d-1) fluxes for the 6

study periods, together with the corresponding average residual

circulation schemes, are summarized in Figures 4 and 5.

In July 2002, Pg decreased from 0.94 ± 0.04 g N m-2 d-1 during the strong

summer upwelling to 0.47 ± 0.03 g N m-2 d-1 during the weak summer

upwelling, because of the progressive reduction of the entry of nutrient

salts. On the other hand, R increased from 13% to 49% of Pg in the photic

layer. Consequently, the net microbial community production, NCP (= Pg

– R) reduced in parallel with the decrease of Pg, from 0.83±0.03 g N m-2 d-1

during the strong summer upwelling to 0.24±0.02 g N m-2 d-1 during the

weak summer upwelling. The resulting average Pg in July was 0.71 g N m-

2 d-1, 27% being respired in the photic zone and 14% in the aphotic zone.

Therefore 0.42 ± 0.03 g N m-2 d-1 produced in the Ría de Vigo in July were

Page 288: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

270

available for microzooplankton grazing, transference to higher trophic

levels, vertical export to the sediments of the ría or horizontal export to

the adjacent shelf. The average f-ratio during July was 0.66. This value is

comparable to the f-ratio obtained by Moncoiffeé et al. (2000) in the same

site from April to November 1991 and also with the value of 0.63 obtained

in the nearby Ría de Arousa by Álvarez-Salgado et al. (1996). The f–ratio

decreased from 0.87 for the strong summer upwelling to 0.53 for the

summer relaxation and 0.40, for weak summer upwelling.

Microzooplankton grazing ranged from 0.60 ± 0.03 g N m-2 d-1 (64±8% of

Pg) during the strong summer upwelling to 0.39 ± 0.02 g N m-2 d-1 (83±9%

of Pg) during the weak summer upwelling. These extremely high grazing

rates implied that, despite the high Pg, there was a reduced transference

of phytoplankton biomass to higher trophic levels. At the ½ week time

scale, the highest Pg values occurred during the spin down of the strong

summer upwelling (15-18 Jul), when phytoplankton grows at the expenses

of the previously upwelled nutrients. During this period, the in vitro NCP

was 0.73±0.04 g N m-2 d-1 (77% of Pg). If microzooplankton grazing (G) is

subtracted to NCP, a surplus of 0.13±0.07 g N m-2 d-1 is obtained that is

comparable with the in situ net ecosystem production (NEP) obtained by

Piedracoba et al. (Chapter 5): -0.13±0.04 g N m-2 d-1. Note that the in situ

NEP, as calculated from N-nutrient changes, is <0 when nutrients are

consumed (= organic matter production) and >0 when nutrients are

produced (= organic matter respiration). The in vitro NCP during the

summer relaxation (18-22 Jul) was 0.39 g N m-2 d-1 (54% of Pg), and

subtracting G, the system would be balanced (= -0.03 ±0.05 g N m-2 d-1).

Accordingly, the in situ NEP was also balanced (-0.05±0.01 g N m-2 d-1;

Piedracoba et al. Chapter 5).

Page 289: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

271

In vitro NCP during the summer weak upwelling (22-26 Jul) was 0.14 g

N m-2 d-1 (30 % of Pg). In this case, Pg was not enough to support

microzooplankton grazing (NCP - G = -0.25 ±0.05 g N m-2 d-1). On the

contrary, the in situ NEP was -0.16±0.03 g N m-2 d-1 (Piedracoba et al.

Chapter 5), which was equivalent to the in vitro NCP but differed

completely when G is subtracted.

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.34±0.05

1.9±0.1

( )0.10±0.010.5±0.03

R

SEDIMENT

Dep0.24±0.06

1.4±0.1

G0.60±0.03

3.3±0.3(64%)

18/07 – 22/07 22/07 – 26/07

(10%)

1.2 Km/d

-0.9 Km/d

-4.0 Km/d

3.0 Km/d

-3.6 Km/d

2.7 Km/d

15/07 – 18/07

NEP0.83±0.03

4.5±0.1

R0.11±0.01

0.7±0.1(13%)(87%)

(o-i)NT(o-i)NT

-0.29±0.02 -0.30±0.01

(27%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.14±0.04

1.0±0.1

( )0.10±0.010.6±0.02

R

SEDIMENT

Dep0.04±0.05

0.4±0.1

G0.42±0.02

2.3±0.2(58%)

(15%)

(o-i)NT0.10±0.01

(10%)

NEP R0.49±0.02

2.8±0.1(69%)

0.23±0.021.2±0.03

(31%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.1+?0.6+?

( )0.10±0.010.6±0.02

R

SEDIMENT

Dep0.0+?

G0.39±0.02

2.2±0.2(81%)

(22%)

(0-8%)

NEP R0.24±0.02

1.4±0.1(51%)

0.23±0.011.3±0.04

(49%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.34±0.05

1.9±0.1

( )0.10±0.010.5±0.03

R

SEDIMENT

Dep0.24±0.06

1.4±0.1

G0.60±0.03

3.3±0.3(64%)

18/07 – 22/07 22/07 – 26/07

(10%)

1.2 Km/d

-0.9 Km/d

-4.0 Km/d

3.0 Km/d

-3.6 Km/d

2.7 Km/d

15/07 – 18/07

NEP0.83±0.03

4.5±0.1

R0.11±0.01

0.7±0.1(13%)(87%)

NEP0.83±0.03

4.5±0.1

R0.11±0.01

0.7±0.1(13%)(87%)

(o-i)NT(o-i)NT

-0.29±0.02 -0.30±0.01

(27%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.14±0.04

1.0±0.1

( )0.10±0.010.6±0.02

R

SEDIMENT

Dep0.04±0.05

0.4±0.1

G0.42±0.02

2.3±0.2(58%)

G0.42±0.02

2.3±0.2(58%)

(15%)

(o-i)NT0.10±0.01

(10%)

NEP R0.49±0.02

2.8±0.1(69%)

0.23±0.021.2±0.03

(31%)

NEP R0.49±0.02

2.8±0.1(69%)

0.23±0.021.2±0.03

(31%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.1+?0.6+?

( )0.10±0.010.6±0.02

R

SEDIMENT

Dep0.0+?

G0.39±0.02

2.2±0.2(81%)

G0.39±0.02

2.2±0.2(81%)

(22%)

(0-8%)

NEP R0.24±0.02

1.4±0.1(51%)

0.23±0.011.3±0.04

(49%)

NEP R0.24±0.02

1.4±0.1(51%)

0.23±0.011.3±0.04

(49%)

∆(V·NT)/ ∆t 0.16±0.02

∆(V·NT)/ ∆t -0.14±0.01

∆(V·NT)/ ∆t 0.12±0.02

Figure 4. Upper panel: Nitrogen (in bold) and carbon (in cursive) fluxes in g N m-2 d-1 and g C m-2 d-1, from in vitro measurements during July 2002. NCP, net community production; R, community respiration; G, microzooplankton grazing; Sed, sedimentation; Dep, deposition; %, fluxes referred to Pg, the gross primary production; (o – i) T, outputs minus inputs of NT; ∆(V·NT)/∆t, accumulation of NT (from Piedracoba et al. Chapter 5). Lower panel: average surface and bottom residual currents (km/d).

Sedimentation also decreased in parallel to the reduction of Pg, from

0.34 ± 0.05 g N m-2 d-1 during the strong summer upwelling to 0.14 ± 0.04

g N m-2 d-1 during the subsequent relaxation. Sedimentation during the

weak summer upwelling should be at least 0.1 g N m-2 d-1 to balance the

respiration of the aphotic layer. It is clear that phytoplankton production

Page 290: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

272

is not able to support the respiration, microzooplankton grazing and

sedimentation rates measured during July, suggesting that i)

sedimentation of microzooplankton occurs; and/or ii) the sediment traps

are collecting a material transported either from the inner ría or from the

shelf (note that the sinking time was between 2 and 8 days). The aphotic

layer respiration was nearly constant in July (0.1 ± 0.01 g N m-2 d-1 or 0.5-

0.6 ± 0.1 g C m-2 d-1) but it increased from 10%of Pg during the strong

summer upwelling to 22% of Pg during the weak summer upwelling. The

balance of sedimentation minus aphotic layer respiration, Dep (= Sed – R),

represents the biogenic material deposited over the pelagic sediments of

the ría. Dep decreased with Pg and sedimentation, and also with the

increase of %R/Pg in the aphotic layer. The aphotic respiration was

independent of the organic matter that sedimented because it was nearly

constant during all the July period.

In September 2002, Pg increased from 0.10 ± 0.02g N m-2 d-1 during the

initial downwelling to 0.28 ± 0.03 g N m-2 d-1 during the final upwelling

period. R also decreased from 83% to 56% of Pg. Consequently, the in vitro

NCP was extremely low during September, ranging from 0.01± 0.01 to

0.14±0.02 g N m-2 d-1. The average f ratio was –0.1. Globally in September

R ≥ NCP and most of Pg was respired in the euphotic layer by the

phytoplankton community.

The low productivity is the reason behind the low microzooplankton

grazing rates compared with July. Anyway, microzooplankton grazed

from 115±32% to 44±12% of Pg throughout the study period, suggesting

that phytoplankton either previously accumulated in the study site or

advected for the inner ría or the shelf are necessary to support

microzooplankton grazing.

Page 291: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

273

The lowest Pg values occurred during the autumn downwelling period

(17-19 Sept) when R exceeded Pg, i.e. the system was heterotrophic (NCP

= -0.10 g N m-2 d-1). If G is subtracted, the deficit was higher (NCP – G = -

0.21 g N m-2 d-1). In contrast, the in situ NEP (Piedracoba et al. Chapter 5)

was -0.02±0.01 g N m-2 d-1, suggesting that the study system was nearly

balanced. In this case, respiration was supported by biogenic materials

accumulated during productive upwelling events and fresh organic

matter imported from the adjacent shelf, as a consequence of the reversal

of the 2D residual circulation pattern as Álvarez-Salgado et al. (1996,

2001), Rosón et al. (1999), Gago et al. (2003) showed previously.

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.15±0.03

1.1±0.1

( )

0.11±0.010.7±0.02

R

SEDIMENT

Dep0.04±0.040.4±0.1

G0.11±0.010.7±0.1(115%)

(117%)

NEP0.01±0.010.1±0.03

R0.09±0.01

0.5±0.03(83%)(17%)

(o-i)NT(o-i)NT

0.12±0.01 -0.10±0.01

(66%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.12 + ?0.8 + ?

( )

0.12±0.010.7±0.02

R

SEDIMENT

Dep0.1+?

G0.17±0.02

1.1±0.1(60%)

(39%)

(o-i)NT0.02±0.01

NEP R0.14±0.020.9±0.04(51%)

0.14±0.010.9±0.03

(49%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.17±0.04

1±0.1

( )

0.11±0.010.7±0.01

R

SEDIMENT

Dep0.06±0.050.3±0.1

G0.11±0.020.7±0.1(44%)

(44%)

(18%)

NEP R0.12±0.020.7±0.04(44%)

0.14±0.010.9±0.03

(56%)

17/09 – 19/09 19/09 – 23/09 23/09 – 26/09

0.5 Km/d

-0.6 Km/d

3.3 Km/d

-3.7 Km/d

-3.4 Km/d

2.5 Km/d

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.15±0.03

1.1±0.1

( )

0.11±0.010.7±0.02

R

SEDIMENT

Dep0.04±0.040.4±0.1

G0.11±0.010.7±0.1(115%)

G0.11±0.010.7±0.1(115%)

(117%)

NEP0.01±0.010.1±0.03

R0.09±0.01

0.5±0.03(83%)(17%)

NEP0.01±0.010.1±0.03

R0.09±0.01

0.5±0.03(83%)(17%)

(o-i)NT(o-i)NT

0.12±0.01 -0.10±0.01

(66%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.12 + ?0.8 + ?

( )

0.12±0.010.7±0.02

R

SEDIMENT

Dep0.1+?

G0.17±0.02

1.1±0.1(60%)

G0.17±0.02

1.1±0.1(60%)

(39%)

(o-i)NT0.02±0.01

NEP R0.14±0.020.9±0.04(51%)

0.14±0.010.9±0.03

(49%)

NEP R0.14±0.020.9±0.04(51%)

0.14±0.010.9±0.03

(49%)

(44% )

0.0±0.1

(0.0% )( )( )

Sed0.17±0.04

1±0.1

( )

0.11±0.010.7±0.01

R

SEDIMENT

Dep0.06±0.050.3±0.1

G0.11±0.020.7±0.1(44%)

G0.11±0.020.7±0.1(44%)

(44%)

(18%)

NEP R0.12±0.020.7±0.04(44%)

0.14±0.010.9±0.03

(56%)

NEP R0.12±0.020.7±0.04(44%)

0.14±0.010.9±0.03

(56%)

17/09 – 19/09 19/09 – 23/09 23/09 – 26/09

0.5 Km/d

-0.6 Km/d

3.3 Km/d

-3.7 Km/d

-3.4 Km/d

2.5 Km/d

∆(V·NT)/ ∆t -0.14±0.01

∆(V·NT)/ ∆t 0.05±0.01

∆(V·NT)/ ∆t 0.07±0.01

Figure 5. Upper panel: Nitrogen (in bold) and carbon (in cursive) fluxes in g N m-2 d-1 and g C m-2 d-1, from in vitro measurements during September 2002. NCP, net community production; R, community respiration; G, microzooplankton grazing; Sed, sedimentation; Dep, deposition; %, fluxes referred to Pg, the gross primary production; (o – i) T, outputs minus inputs of NT; ∆(V·NT)/∆t, accumulation of NT (from Piedracoba et al. Chapter 5). Lower panel: average surface and bottom residual currents (km/d).

Page 292: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

274

During the autumn relaxation period, in vitro NCP was 0.02±0.03 g N

m-2 d-1 (7 % of Pg). Considering that G was 0.17±0.02 g N m-2 d-1, NCP was

not enough to support the microheterotrophs community. The in situ NEP

(0.07±0.01 g N m-2 d-1.) was comparable with the in vitro approach.

In vitro measurements during the weak autumn upwelling showed a

NCP of 0.01± 0.03 g N m-2 d-1, i.e. a nearly balanced system unable to

support the microzooplankton activity. The in situ NEP also showed that

the system was balanced.

Despite the low productivity, sedimentation rates in September were

comparable to July. Again, sinking times between 5 and 8 d suggest that

the sediment trap is not collecting the biogenic material produced during

the collection time but the previously accumulated or transported either

from the inner ría or the shelf. The aphotic layer R was nearly constant

during September at 0.11 ± 0.01 g N m-2 d-1 (0.7 ± 0.1 g C m-2 d-1).

Sedimentation during the autumn transition should be at least 0.12 g N m-

2 d-1 (0.8 g C m-2 d-1) to balance the respiration of the aphotic layer. Dep

fluxes were small in comparison with July.

Considering all sampling dates together, the average Pg coincided with

the seasonal mean of 2.5 ± 2.8 g C m-2 d-1 estimated for the rías during the

upwelling season (Fraga 1976, Bode and Varela 1998, Tilstone et al. 1999,

Moncoiffé et al. 2000, Teira et al. 2003, Varela et al. 2004, 2005, Lorenzo et

al. 2005). The average R in the photic and aphotic layer were also

comparable with the seasonal mean estimates of 43% and 25% of Pg

respectively, obtained by Moncoiffé et al. (2000) in the same site during

the upwelling season of 1991.

The results show a close coupling between microzooplankton grazing

and phytoplankton growth (g ~ µ) between 15 and 22 July. This coupling

Page 293: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

275

might not be expected during upwelling conditions when the

phytoplankton community tends to be dominated by large cells,

particularly diatoms (Cushing 1989). However the coupling was equally

found in the shelf during an upwelling/relaxation event (Fileman et al.

2001), where despite the increasing diatom presence, pico- and

nanoflagellates dominated surface phytoplankton populations (Joint et al.

2001) and these supported the community of nano- and microplankton.

Fileman et al. (2001) affirmed that in situations where close coupling

between microzooplankton herbivory and phytoplankton growth was

observed, microzooplankton grazing play a key role in the rapid recycling

of carbon within the microbial loop. The high microzooplankton grazing

rates measured in the photic layer contributes to keep most of the

phytoplankton production in the microbial loop trophic level, constituted

by the microheterotrophs (bacteria and protozoa) and the autotrophic pro-

and eukaryotic unicellular organisms (Sherr and Sherr, 1998). Trophic

interactions within the microbial loop implied a lower efficiency in the

transference of organic matter to the metazoan food web.

The only studies of vertical biogenic particle flux in this region were

the measurements carried out in the Ría de Pontevedra by Varela et al.

(2004) and the shelf and slope measurements by Olli et al. (2001). Varela et

al. (2004) obtained sedimentation fluxes in the same range of magnitude

(0.5 to 1.8 g C m-2 d-1) than in the Ría de Vigo. The sedimentation fluxes

measured in the Ría de Pontevedra were quantitatively related to the

hydrography: a negative relationship was obtained between

sedimentation and horizontal advection. The reduced number of study

cases in our work limits prevented us to look for any quantitative

relationship between the sedimentation and hydrography of the Ría de

Page 294: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

276

Vigo and only a qualitative approach can be made. In July, the nutrient

entry during strong summer upwelling allowed phytoplankton growth.

The subsequent relaxation, caused phytoplankton at the surface on 18 July

to sink to a subsurface maximum of Chl a, fucoxanthin and Chl c3 on 22

July. However, during the weak summer upwelling, hydrodynamic

conditions favoured phytoplankton in the photic layer to flow out of the

ría by the outgoing surface current.

In September, despite the low productivity, sedimentation rates were

similar to July. The low nutrient levels of the photic layer and the

dominant downwelling conditions during the study period are likely the

reasons behind this high sedimentation. Aggregates of phytoplankton

cells usually occur when nutrients are depleted, producing high sinking

rates (Riebesell 1989, Álvarez-Salgado 2005). Large sedimentation rates

have been reported in the literature in relation to nutrient depletion in the

photic layer by Smayda (1970), Waite et al. (1992) and Richardson and

Cullen (1995). Pigment distributions indicate that a dinoflagellates bloom

developed during the autumn downwelling, which sank to the 25% PAR

during the autumn relaxation. The subsurface maximum was dispersed

by the surface outgoing current developed during the subsequent weak

autumn upwelling. On the contrary, Cryptophyceae sp., marked by the

surface alloxantin maximum during the autumn downwelling, sank to the

bottom during the autumn relaxation and was reintroduced in the ría by

the bottom flow during the autumn upwelling. In these sense, the ability

of dinoflagellates for vertical migration (Figueiras et al. 1994, Fermín et al.

1996) could keep them at mid water depths during the autumn relaxation

unlike cryptophytes sank to the bottom.

Page 295: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

277

Olli et al. (2001) studied the vertical flux and composition of settling

biogenic matter during a lagrangian experiment off the NW Spanish

continental margin. They obtained POC fluxes of 90-240 mg C m-2 d-1 over

the shelf and 50-180 mg C m-2 d-1 in the adjacent ocean. Although the

magnitude of these sedimentation rates was lower than the fluxes

obtained in the ría, they represented 14-26% of the integrated primary

production per day on the shelf and 25-42% on the slope, i.e. similar

percentages than during the upwelling/relaxation cycle of July 2002 in the

Ría de Vigo.

The geochemical and microbial approaches were in agreement in three

of the six periods analyzed in this study. The discrepancies between them

intensified due to the high grazing fluxes, which revealed that the

immediate fate of NCP was the microbial loop rather than being

transferred to the metazoans. The imbalance between NCP and G

suggests that the system needs to import allochthonous organic matter to

support the microheterotrophs. In July, the hydrodynamics (positive

residual circulation) indicates that the organic matter is basically imported

from the inner ría. In contrast, the dominant hydrodynamics in September

(negative residual circulation) means that the source of organic matter to

support the metabolic requirements arrived from the shelf.

Several factors contribute to enlarge the differences between the

geochemical and in situ balances. The in situ balance is representative for

a station (five depths), while the geochemical balance considers a

complete section of the ría. In addition, the incubations preformed to

analyse the metabolic state of the system implies are affected by the so

called “bottle effects”. During the incubation time (24 hours) the initial

conditions are modified and artefacts associated with isolating a portion

Page 296: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

278

of the community in containers can appear: reduced turbulence, unnatural

light field and altered grazer communities. In addition, environmental

changes during the incubation time can not be reflected with an in vitro

method, a relevant issue in a transient marine ecosystem such as the Ría

de Vigo. Furthermore, incubations may give precise measurements for

each component, but with method-dependent biases. If such biases occur,

then the summed measurement will not accurately represent the net

metabolism (Smith and Hollibaugh 1993). On the other hand, the

processes that occur in incubation bottles filled in a single site do not catch

the different environments enclosed by the middle section of the ría (e.g.

the culture of mussels on hanging ropes or the sedimentary basin of San

Simon Bay). Finally, the high microzooplankton grazing is not coherent

with respiration, maybe because of the lag time between ingestion and

respiration processes is likely to occur in the microheterotrophs

community. Besides, the grazing rates measure the instantaneous changes

in chlorophyll or carbon concentrations and the respiration rates are

influenced by previous processes.

The geochemical approach avoids most of the problems associated

with in vitro methods and, although box-model estimations involve large

potential errors in the estimation of the NEP of N–nutrients, DON and

PON, it is a unique method to measure the ecosystem metabolism directly

(Smith and Hollibaugh 1997). Although box model estimations are surely

less accurate than in vitro techniques, the resultant values are ready for

direct interpretation at the ecosystem level. The deficits of the geochemical

approach are linked to the fact that the specific processes that occur in the

study system can not be identified, i.e. only the net result of the processes

taking place inside the box is known. However, accumulation and losses,

Page 297: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

In vitro approach to the physical–biological coupling in the Ría de Vigo

279

a relevant component of the metabolic balance of any transient ecosystem

(Álvarez-Salgado et al. 2001, Piedracoba et al. Chapter 5) omitted by the in

vitro approach, can be evaluated.

Considering the advantages of the in vitro and in situ approaches, a

combination of both methods would be the best option to identify the

importance of the processes that occur in the Ría de Vigo and to define the

metabolic state of the system.

Page 298: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 6

280

5. Conclusions

Measurements of pelagic community production and respiration were

performed in the middle section of the Ría the Vigo during two periods,

July and September 2002. For first time, microzooplankton grazing and

sedimentation rates were measured concomitantly. The high grazing rates

questions the capability of the ría to transfer organic matter to higher

trophic levels despite its high productivity and provokes disagreements

between the in vitro approach in this paper and the in situ approach of

Piedracoba et al. (Chapter 5) to asses the metabolic status of the ría. It is

concluded that both methods are complementary and the simultaneous

application allowed us to obtain a better knowledge of how the system

operates. The in situ approach is adequate to define the metabolic state of

the system and the in vitro approach to know the relative importance of

the variety of metabolic processes that occur in the Ría de Vigo.

Page 299: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 7:

General Conclusions

Page 300: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 301: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

General Conclusions

283

Gran parte del conocimiento sobre la hidrodinámica de las Rías Baixas

se ha desarrollado en la última década a partir de modelos inversos

basados en datos hidrográficos, debido tanto a la escasez de medidas

directas de corrientes como a su insuficiente cobertura espacio-temporal.

En esta memoria se ha analizado la variabilidad hidrodinámica del

sistema de afloramiento costero de la Ría de Vigo a corta escala espacio-

temporal. Para ello se ha contado con un registro continuo de corrientes

obtenido a partir del fondeo de un correntímetro Doppler en el segmento

central de la Ría de Vigo durante cuatro escenarios hidrográficos

característicos.

Asimismo, las medidas de corrientes registradas en el centro de la ría

se han complementado con las obtenidas con un perfilador acústico

doppler montando en el casco del buque oceanográfico Mytilus, y con las

simulaciones del modelo numérico HAMSOM, para ofrecer una

descripción realista de la asimetría lateral de la circulación residual así

como de sus fuentes de variabilidad en la Ría de Vigo.

Desde el punto de vista hidrodinámico, en relación con los

forzamientos dominantes, puede concluirse lo siguiente:

1. La circulación residual en la parte central de la Ría de Vigo bajo

condiciones hidrográficas y meteorológicas contrastadas se

distribuye normalmente en dos capas. Este tipo de circulación

había sido asumida como hipótesis previamente para poder

ejecutar modelos inversos, y se había obtenido en los resutltados

de modelos numéricos. Por otra parte, el registro continuo de

corrientes nos ha permitido identificar periodos de circulación

cortos en tres capas, que habitualmente se producen en las

Page 302: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 7

284

transiciones de circulación residual positiva a negativa y

viceversa.

2. El viento de plataforma explica más del 65% de la variabilidad de

la circulación residual, mientras que el intercambio de calor con la

atmósfera y los aportes continentales explican menos del 25% y el

5%, respectivamente. El viento local no contribuye

significativamente a la variabilidad de la circulación residual de la

ría.

3. La circulación residual responde al viento de plataforma de forma

inmediata. Aunque el retardo de la corriente en relación al viento

se encuentra entre 0 y 2 días, dependiendo de la profundidad y de

la época del año, comúnmente la respuesta hidrodinámica al

viento de plataforma se produce con menos de 1 día. Esta

respuesta casi instantánea está relacionada probablemente con el

pequeño tamaño de ría y con su topografía (canal en forma de

embudo).

4. Se han encontrado diferencias laterales significativas en las

componentes longitudinal y trasversal de la corriente residual en

el segmento central de la Ría de Vigo, así como en los perfiles de

salinidad y temperatura, mostrando que la circulación secundaria

es relevante. Esta circulación secundaria no se explica por efectos

rotacionales (fuerza de Coriolis), sino por la acción combinada del

viento local y de plataforma. El viento de plataforma interactúa

con la topografía de la parte más externa de la ría, definida por la

boca Sur, más ancha y profunda, separada por las Islas Cíes de la

boca Norte, más estrecha y somera. Esta interacción da lugar a un

patrón de circulación residual tridimensional en la parte más

Page 303: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

General Conclusions

285

externa de la ría que puede ser responsable de la aparición de

otras aceleraciones no lineales (advección y fricción) que juegan

un papel importante en la dinámica lateral de la ría,

especialmente en la componente longitudinal de la corriente

residual. El viento local, por otra parte, ejerce su influencia en la

componente transversal de la velocidad residual. Así, la acción

combinada de la componente transversal del viento local y los

efectos rotacionales da lugar a celdas con sentidos de circulación

opuestos entre las capas de superficie y fondo.

Simultáneamente al muestreo hidrodinámico, se realizó un muestreo

hidrográfico y biogeoquímico con el fin de estudiar el impacto de los

procesos dinámicos en los procesos biogeoquímicos. Se determinaron una

serie de variables de estado (oxígeno disuelto, clorofila, sales nutrientes y

materia orgánica disuelta y particulada), así como una serie de tasas

biogeoquímicas (producción primaria, respiración, herbivoría y

sedimentación). La medición de distintas variables de estado en cada

muestreo permitió la construcción de un modelo adaptado a la

estequiometría del sistema en lugar de realizar conversiones

estequiométricas prefijadas. Mediante este modelo estequiométrico, todas

las tasas biogeoquímicas medidas en distintas unidades (oxígeno,

clorofila, carbono) se expresaron en la misma unidad (nitrógeno).

Desde el punto de vista biogeoquímico se han estudiado tres escenarios

hidrográficos. Así en el mes de Febrero de 2002 se siguió la iniciación,

desarrollo y finalización de un “bloom” de la diatomea Skeletonema

costatum durante un período de 10 días.

Las conclusiones más relevantes de este trabajo fueron:

Page 304: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 7

286

5. La variabilidad de las condiciones hidrográficas del Noroeste de

la Península Ibérica, caracterizada por la sucesión de episodios de

tensión/relajación del viento costero, permitió el desarrollo de un

“bloom” inusual de fitoplancton durante el invierno, que sería

más típico de la primavera. El inicio, desarrollo y terminación del

“bloom” de diatomeas estuvo marcado por la sucesión de

condiciones de afloramiento, relajación y hundimiento. El

“bloom”, durante el cual la producción primaria bruta alcanzó

valores de 8 g C m–2 d–1, se produjo en el período de relajación de

los vientos costeros de 3-4 días y fue promovido por los nutrientes

aflorados durante el período anterior. En la terminación del

“bloom”, el proceso físico de hundimiento fue determinante en

cuanto al destino de la materia orgánica producida previamente,

bien directamente a través de la advección vertical dirigida hacia

la capa inferior (14% del flujo vertical total), bien indirectamente a

través de la sedimentación (86% del flujo vertical total). La

agregación celular y posterior sedimentación se promueve en

condiciones de ausencia de nutrientes en la capa fótica, típicas de

situaciones de hundimiento.

6. El destino principal de la materia orgánica producida durante el

“bloom” primaveral fue el ecosistema bentónico, en lugar de ser

consumida en el ecosistema pelágico. La importancia de la

sedimentación inferida a partir del balance biogeoquímico

aplicado al periodo invernal nos condujo a medir las tasas de

sedimentación durante los periodos de Julio y Septiembre de 2002

por primera vez en la Ría de Vigo. El inusual comportamiento del

periodo invernal, típico de condiciones primaverales, nos condujo

Page 305: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

General Conclusions

287

a no considerar un trabajo específico del mes de Abril de 2002. El

nuevo enfoque consistió en comparar dos escenarios

oceanográficos, el afloramiento/relajación estival (Julio 2002) con

la proliferación otoñal (Septiembre 2002), bajo dos perspectivas

biogequímicas un balance in situ versus un balance in vitro.

Así, las conclusiones más relevantes correspondientes a la comparación

de los dos escenarios oceanográficos son:

7. Se ha evaluado por primera vez el balance geoquímico de la Ría

de Vigo con datos de corrientes medidos (en lugar de inferir los

flujos a partir de variables hidrográficas) teniendo en cuenta la

variación lateral de la concentración de nutrientes, materia

orgánica disuelta y particulada, así como de la corriente residual

que conjuntamente originan una distribución no homogénea de

los flujos de compuestos orgánicos e inorgánicos a lo largo de la

sección central de la Ría de Vigo.

8. Se ha evaluado la importancia relativa de la convección vertical y

la mezcla en la fertilización de la capa fótica. La mezcla puede ser

el mecanismo más importante de fertilización en condiciones de

afloramiento intenso (el 15 de julio de 2002, Kz/∆z y Vz fueron

8,2±0,2 m d-1 y 3,7±0,5 m d-1, respectivamente).

9. La Ría de Vigo presentó un comportamiento metabólico

autotrófico en Julio y de balance metabólico en Septiembre. La

ventaja de los balances in situ es que, aplicados a una resolución

espacio temporal adecuada, permiten tener en cuenta la

acumulación neta en un sistema no estacionario como la Ría de

Vigo, en el que la acumulación domina transitoriamente a la

exportación de la materia orgánica.

Page 306: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 7

288

10. Se ha estudiado el acoplamiento de la producción primaria y la

respiración a escalas estacionales y episódicas a partir de

experimentos in vitro, ampliando su resolución vertical respecto a

trabajos anteriores. Se ha medido específicamente la respiración

en la zona fótica y afótica. La producción primaria bruta en Julio

de 2002 fue en promedio de 0,71 g N m-2 d-1, siendo un 27%

respirada en la capa fótica y un 14% en la capa afótica. En

Septiembre de 2002 la producción primaria bruta fue en promedio

de 0,22 g N m-2 d-1, siendo un 63% respirada en la capa fótica y un

67% en la capa afótica.

11. Se han medido por primera vez tasas de herbivoría y de

sedimentación en la Ría de Vigo. La relevancia de ambos flujos

cuestiona la idea de considerar la ría como un sistema con elevada

capacidad para transferir materia orgánica a niveles tróficos

superiores. A lo largo del muestreo de Julio de 2002 el

microzooplankton (ciliados y flagelados) consumió el 68% de la

producción primaria, mientras que la sedimentación representó el

33% de la producción primaria. Durante el muestreo de

Septiembre de 2002 la herbivoría y la sedimentación

representaron el 38% y el 107% de la producción primaria,

respectivamente. A diferencia de Julio y Septiembre de 2002, en

Febrero de 2002 el consumo de materia orgánica por parte de los

microheterótrofos (29% de la producción primaria) tiene menor

importancia relativa en comparación con el transporte vertical por

sedimentación (48% de la producción primaria).

12. La controversia entre el carácter más o menos autotrófico de la ría

se pone de manifiesto cuando las dos aproximaciones metabólicas

Page 307: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

General Conclusions

289

estudiadas se enfrentan: el balance in situ versus el balance in vitro.

El balance in situ es más apropiado para definir el estado

metabólico del sistema debido a que no se modifican las

condiciones ambientales y se tiene en cuenta las características no

estacionarias del sistema. El balance in vitro complementa al

balance in situ porque permite identificar la importancia relativa

de los procesos metabólicos que ocurren en la ría.

13. Es necesario ser cauto a la hora de comparar ambos balances, ya

que los balances in situ e in vitro se aplican a escalas espaciales

(km3 frente a cm3) y temporales (3-4 días frente a 1 día)

completamente distintas, además del “efecto botella” en las

incubaciones y el uso de colectores en las trampas de

sedimentación. Por otra parte, la magnitud de las tasas de

sedimentación implica que una proporción elevada de la

producción primaria alcanza los sedimentos. Sin embargo, el

tiempo de sedimentación es suficientemente alto como para que

la fracción sedimentada no sea representativa de la producción

primaria medida simultáneamente en la capa fótica.

Page 308: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 7

290

Page 309: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8:

Literature Cited

Page 310: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade
Page 311: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

293

Agsterberg R, Wieringa J. 1989. Mesoscale terrain roughness mapping of

the Netherlands. In: Royal Netherlands Meteorological Institute

(Ed.), Technical Report TR–115.

Álvarez I, deCastro M, Prego R, Gomez–Gesteira R. 2003. Hydrographic

characterization of a winter–upwelling event in the Ria of

Pontevedra (NW Spain). Estuar. Coast. Shelf Sci. 56, 869–876.

Álvarez–Salgado XA, Rosón G, Pérez FF, Pazos Y. 1993. Hydrographic

variability off the Rías Baixas (NW Spain) during the upwelling

season. J. Geophys. Res. 98, 14447–14455.

Álvarez–Salgado XA, Rosón G, Pérez FF, Figueiras FG, Pazos Y. 1996a.

Nitrogen cycling in an estuarine upwelling system, the Ría de

Arousa (NW Spain). I. Short–time–scale patterns of hydrodynamic

and biogeochemical circulation. Mar. Ecol. Prog. Ser. 135, 259–273.

Álvarez–Salgado XA, Rosón G, Pérez FF, Figueiras FG, Ríos AF. 1996b.

Nitrogen cycling in an estuarine upwelling system: the Ría de

Arousa (NW, Spain). II: Spatial differences in the short–time scale

evolution of fluxes and net budgets of nitrogen species. Mar. Ecol.

Prog. Ser. 135, 275–288.

Álvarez–Salgado XA, Castro CG, Pérez FF, Fraga F. 1997. Nutrient

mineralization patterns in shelf waters of the Western Iberian

upwelling. Cont. Shelf. Res. 17, 1247–1270.

Álvarez–Salgado XA, Figueras FG, Villarino ML, Pazos Y. 1998a.

Hydrodynamic and chemical conditions during onset of a red-tide

assemblage in an estuarine upwelling ecosystem. Mar. Biol. 130, 509-

519.

Page 312: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

294

Álvarez–Salgado XA, Miller AEJ. 1998. Simultaneous determination of

dissolved organic carbon and total dissolved nitrogen in seawater

by high temperature catalytic oxidation: conditions for precise

shipboard measurements. Mar. Chem. 62, 325–333.

Álvarez–Salgado XA, Gago J, Míguez BM, Gilcoto M, Pérez FF. 2000.

Surface waters of the NW Iberian Margin: Upwelling on the shelf

versus outwelling of upwelled waters from the Rías Baixas. Estuar.

Coast. Shelf Sci. 51, 821–837.

Álvarez–Salgado XA, Gago J, Míguez BM, Pérez FF. 2001. Net ecosystem

production of dissolved organic carbon in a coastal upwelling

system: The Ría de Vigo, Iberian margin of the North Atlantic.

Limnol. Oceanogr. 46, 135– 147.

Álvarez–Salgado XA, Beloso S, Joint I, Nogueira E, Chou L, Pérez FF,

Groom S, Cabanas JM, Rees AP, Elskens M. 2002. New Production

of the NW Iberian Shelf during the Upwelling Season over the

period 1982–1999. Deep–Sea Res. I 49, 1725–1739.

Álvarez–Salgado XA, Figueiras FG, Pérez FF, Groom S, Nogueira E,

Borges AV, Chou L, Castro CG, Moncoiffé G, Ríos AF, Miller AEJ,

Frankignoulle M, Savidge G, Wollast R. 2003. The Portugal Coastal

Counter Current off NW Spain: New Insights on its Biogeochemical

Variability. Prog. Oceanogr. 56, 281–321.

Álvarez-Salgado XA, Gilcoto M. 2004. Inferring nitrification rates with an

inverse method in a coastal upwelling system, Ría de Vigo (NW

Spain). Mar. Ecol. Prog. Ser. 276, 3–17

Page 313: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

295

Álvarez–Salgado XA, Nieto–Cid M, Piedracoba S, Crespo BG, Gago J,

Brea S, Teixeira IG, Figueiras FG, Garrido JL, Rosón G, Castro CG,

Gilcoto M. 2005. Origin and fate of a bloom of Skeletonema costatum

during a winter upwelling/downwelling sequence in the Ría de

Vigo (NW Spain). J. Mar. Res. 63, 1127–1149. Chapter 4 of this thesis.

Álvarez–Salgado XA , Fraga F. As Rías. Consello da Cultura Galega (in

press).

Álvarez-Salgado XA, Nieto-Cid M, Gago J, Brea S, Castro CG, Doval MD,

Pérez FF. 2006. Stoichiometry of the mineralization of dissolved,

suspended and sinking organic matter in the NW Iberian upwelling.

J. Geophys. Res. 111(C07017), doi:10.1029/2004JC002473.

Álvarez–Salgado XA, Borges AV, Figueiras FG, Chou L. Carbon and

nutrient fluxes and budget in the rías of the NW Iberian upwelling

system, chapter 2.6 in Carbon and Nutrient Fluxes in Continental

Margins: A Global Synthesis, Liu K–K, Atkinson L, Quinones R,

Talaue–McManus L (Eds), Springer–Verlag, New York (in press).

Anderson LA. 1995. On the hydrogen and oxygen content of marine

phytoplankton. Deep–Sea Res. I 42, 1675–1680.

Asplin L. 1995. Examination of local circulation in a wide, stratified fjord

including exchange of water with the adjacent ocean, due to

constant local up fjord wind. In: Ecology of Fjords and Coastal

Waters. Skjoldal HR (Ed). Elsevier, New York, 177–194.

Arístegui J, Álvarez–Salgado XA, Barton ED, Figueiras FG, Hernández–

León S, Roy C,Santos AMP. 2006. Oceanography and Fisheries of the

Canary Current/Iberian Region of the Eastern North Atlantic. In:

The Sea, vol 14. Brink KH, Robinson AR (Eds). John Wiley & Sons,

New York, 877–931.

Page 314: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

296

Armstrong FAJ, Williams PM, Strickland JDH. 1966. Photo–oxidation of

organic matter in sea water by ultraviolet radiation, analytical and

other applications. Nature 211, 481–483.

Azam F, Fenchel T, Field Jg, Meyer-Reil RA, Thingstad F. 1983. The

cycling of organic matter by bacterioplanktonic pelagic marine

systems: Microenvironmental considerations. In: Flows of Energy

and Materials in Marine Ecosystems: Theory and Practice. NATO

Conference Series 4, Marine Sciences, Fashman MJR (Ed). Plenum

Press, New York, 345-360.

Bakun A, Nelson CS. 1991. The seasonal cycle of wind–stress curl in

subtropical eastern boundary current regions, J. Phys. Oceanogr. 21,

1815–1834.

Bienfang PK. 1981. Sinking rates of heterogeneous, temperate

phytoplankton populations. J. Plankton Res. 3, 235–253.

Blanton JO, Tenore K.R., Castillejo FF de, Atkinson LP, Schwing FB, Lavín

A. 1987. The relationship of upwelling to mussel production in the

rías on the Western coast of Spain. J. Mar. Res. 45, 497–511.

Blasco D, Estrada M, Jones B. 1980. Relations between the phytoplankton

distribution and composition and the hydrography in the upwelling

region near Cabo Corbeiro. Deep–Sea Res. 27, 799–821.

Bode A, Varela M. 1998. Primary production and phytoplankton in three

Galician Rías Altas (NW Spain): seasonal and spatial variability. Sci.

Mar. 62, 319–330.

Bode A, Varela MM, Teira E, Fernández E, González N, Varela M. 2004.

Planktonic carbon and nitrogen cycling off Northwest Spain:

variations in production of particulate and dissolved pools. Aquat.

Microb. Ecol. 37, 95–107.

Page 315: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

297

Bode A, González N, Rodríguez C, Varela M, Varela MM. 2005. Seasonal

variability of plankton blooms in the Ria de Ferrol (NW Spain): I.

Nutrient concentrations and nitrogen uptake rates. Estuar. Coast.

Shelf Sci. 63, 269–284.

Brink KH, Jones BH, Van Leer JC, Mooers CNK, Stuart DW, Stevenson

MR, Dugdale RC, Heburn GW. 1980. Physical and biological

structure and variability in an upwelling center off Peru near 15°S

during March 1977. In: Coastal Upwelling. Richards FA (Ed).

American Geophysical Union, Washington DC, pp. 473–495.

Broecker WS, Peng TS. 1982. Tracers in the Sea. Lamont-Doherty

Geological Observatory, Palisades, New York, 690 pp.

Bruno M, Rico J, Rosario JJA, Ruíz–Cañavate A, Mañanes R. 1998. A

simple method for tidal current prediction in short length

convergent channels. An application to the Vigo and Pontevedra

estuaries, Int. Hydrog. Rev. 75, 135–152.

Brzezinski MA. 1985. The Si:C:N ratio of marine diatoms: interspecific

variability and the effect of some environmental variables. J. Phycol.

21, 347–357.

Buesseler KO. 1991. Do upper-ocean sediment traps provide an accurate

record of particle flux? Nature 353, 420–423.

Cáceres M, Valle–Levinson A, Sepúlveda HH, Holderied K. 2002.

Transverse variability of flow and density in a Chilean fjord. Cont.

Shelf Res. 22, 1683–1698.

Cáceres M, Valle–Levinson A. 2004. Transverse variability of flow on both

sides of a sill/contraction combination in a fjord–like inlet of

southern Chile. Estuar. Coast. Shelf Sci. 60, 325–338.

Page 316: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

298

Calbet A, Landry MR. 2004. Phytoplankton growth, microzooplankton

grazing and carbon cycling in marine systems. Limnol. Oceanogr.

49, 51–57.

Carter, HH, Najarian TO, Pritchard DW, Wilson RE. 1979. The dynamics

of motion in estuaries and other coastal water bodies. Rev. Geophys.

17, 1585–1590.

Castillejo FF, Lavín A. 1982. Contribución al estudio del flujo de agua

entrante y saliente en la Ría de Arosa. Bol. Inst. Esp. Oceanogr. 7,

163–180.

Castro CG. 1997. Caracterización química del agua subsuperficial del

Atlántico Nororiental y su modificación por procesos

biogeoquímicos, Ph.D. Thesis, Univ. of Santiago, 244pp.

Castro CG, Álvarez-Salgado XA, Figueiras FG, Pérez FF, Fraga F. 1997.

Transient hydrographic and chemical conditions affecting

microplankton populations in the coastal transition zone of the

Iberian upwelling system (NW Spain) in September 1986. J. Mar.

Res. 55, 321-352

Cermeño P, Marañón E, Rodríguez J, Fernández E. 2005. Large–sized

phytoplankton sustain higher carbon–specific photosynthesis than

smaller cells in a coastal eutrophic ecosystem. Mar. Ecol. Prog. Ser.

297, 51–60.

Chase J. 1975. Wind–driven circulation in a Spanish estuary. Estuar. Coast.

Mar. Sci. 3, 303–310.

Chen RF, Fry B, Hopkinson CS, Repeta DJ, Peltzer ET. 1996. Dissolved

organic carbon on Georges Bank. Cont. Shelf. Res. 19, 409-420.

Page 317: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

299

Cotner JB, Biddanda BA. 2002. Small players, large role: microbial

influence on biogeochemical processes in pelagic aquatic

ecosystems. Ecosystems 5, 105-121.

Csanady GT. 1973. Wind–induced barotropic motions in long lakes. J.

Phys. Oceanogr. 3, 429–438.

Csanady GT. 1975. Lateral momentum flux in boundary currents. J. Phys.

Oceanogr. 5, 705–717.

Culver ME, Smith Jr WO. 1989. Effects of environmental variation of

sinking rates of marine phytoplankton. J. Phycol. 25, 262–270.

Cusham–Roisin B. 1994. Introduction to Geophysical fluid dynamics.

Precinte Hall, Englewood Cliffs, New Jersey.

Cushing DH. 1989. A difference in structure between ecosystems in

strongly stratified wasters and in those that are only weakly

stratified. J. Plank. Res. 11, 1-13.

del Giorgio PA, Cole JJ, Cimbleris A. 1997. Respiration rates in bacteria

exceed phytoplankton production in unproductive aquatic systems.

Nature 385, 148-151.

del Giorgio PA, Duarte CM. 2002. Respiration in the open ocean. Nature

420, 379-384.

Diz P, Francés G, Costas S, Souto C, Alejo I. 2004. Distribution of benthic

foraminifera in coarse sediments, Ría de Vigo, NW Iberian Margin.

J. Foraminif. Res. 34 (4), 258–275, doi: 10.2113/34.4.258.

Doval MD, Álvarez–Salgado XA, Pérez FF. 1997. DOM in a temperate

embayment affected by coastal upwelling. Mar. Ecol. Prog. Ser. 157,

21–37.

Page 318: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

300

Druon y Le Fèvre. 1999. Sensitivity of a pelagic ecosystem model to

variations of process parameters within a realistic range. J. Mar.

Syst. 19, 1–26.

Duarte C, Agustí S. 1998. The CO2 balance of unproductive aquatic

systems. Science 281, 234-236.

Ducklow HW, Carlson CA, Bates NR, Knap AH, Michaels AF, Jickells T,

Williams PJ Le B, McCave IN. 1995. Dissolved Organic Carbon as a

Component of the Biological Pump in the North Atlantic Ocean.

Phil. Trans. Bio. Sci. 348, 161-167.

Dugdale RC, Goering JJ. 1967. Uptake of new and regenerated forms of

nitrogen in primary production. Limnol. Oceanogr. 12, 196-206.

Dyer KR. 1974. The salt balance in stratified estuaries. Estuar. Coast. Mar.

Sci. 2, 273.

Dyer KR. 1977. Lateral circulation effects in estuaries. In: Estuaries,

Geophysics of the Environment, National Research Council,

National Academy of Science, Washington D.C., pp. 22–29.

Dyer KR. 1997. Estuaries, A Physical Introduction, 2nd Edition. Wiley,

New York.

Edwards ES, Burkill PH, Stelfox CE. 1999. Zooplankton herbivory in the

Arabian Sea during and after the SW Monsoon, 1994. Deep–Sea Res.

II 46, 843–863.

Edwards A, Jones K, Graham JM, Griffiths CR, MacDougall N, Patching J,

Patching JM, Richard JM, Raine R. 1996. Transient coastal upwelling

and water circulation in Bantry Bay, a ria on the South–west Coast

of Ireland. Estuar. Coast. Shelf Sci. 42, 213–230.

Elliot AJ. 1978. Observations of the meteorologically induced circulation in

the Potomac Estuary. Estuar. Coast. Mar. Sci. 6, 285–290.

Page 319: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

301

Eppley RW, Peterson BJ. 1979. Particulate organic matter flux and

planktonic new production in the deep ocean. Nature 282, 677–680.

Estrada M, Blasco D. 1979. Two phases of the phytoplankton community

in the Baja California upwelling. Limnol. Oceanogr. 24, 1065–1080.

Fermín EG, Figueiras FG, Arbones B, Villarino ML. 1996. Short–time scale

development of a Gymnodinium catenatum population in the Ría de

Vigo (NW Spain). J. Phycol. 32, 212–221.

Figueiras FG, Niell FX. 1986. Visibilidad del disco de Secchi, clorofila a y

materia orgánica particulada en las Ría de Pontevedra (NO de

España). Inv. Pesq. 50, 607–637.

Figueiras FG, Niell FX. 1987. Composición del fitoplancton en la rías de

Pontevedra, NO de España. Inv. Pesq. 51, 371–409.

Figueiras FG, Ríos AF. 1993. Phytoplankton succession, red tides and the

hydrographic regime in the Rias Bajas of Galicia. In: Toxic

Phytoplankton blooms in the sea. Smayda TJ, Shimizu Y (Eds).

Elsevier Science Publishers B.V., pp. 239–244.

Figueiras FG, Jones KJ, Mosquera AM, Álvarez–Salgado XA, Edwards A,

MacDougall N. 1994. Red tide assemblage formation in an estuarine

upwelling ecosystem: Ría de Vigo. J. Plankton Res. 16, 857–878.

Figueiras FG, Labarta U, Fernández Reiriz MJ. 2002. Coastal upwelling,

primary production and mussel growth in the Rías Baixas of Galicia.

Hidrob. 484, 121–131.

Fileman E, Burkill P. 2001. The herbivorous impact of microzooplankton

during two short–term Lagrangian experiments off the NW coast of

Galicia in summer 1998. Prog. Oceanogr. 51, 361–383.

Fischer HB. 1972. Mass transport mechanisms in partially stratified

estuaries. J. Fluid Mech. 53, 671–687.

Page 320: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

302

Fiúza, AFG. 1984. Hidrologia e dinâmica das aguas costeiras de Portugal.

PhD thesis, University of Lisbon, 294 pp.

Fraga F. 1976. Fotosíntesis en la Ría de Vigo. Inv. Pesq. 40, 151–167.

Fraga, F. 1981. Upwelling off the Galician coast, Northwest Spain in

Coastal Upwelling, vol. 1, Coastal and Estuarine Science, Richards

FA (Ed), AGU, Washington, D. C.,pp 176–182.

Fraga F, Mouriño C, Manríquez M. 1982. Las masas de agua en la costa de

Galicia: junio–octubre, Res. Exped. Cient. 10, 51–77.

Fraga F, Barton ED, LLinas O. 1985. The concentration of nutrient salts in

"pure" North and South Atlantic Central waters. Inv. Pesq.

Barcelona 5, 25–36.

Fraga F, Pérez FF, Figueiras FG, Ríos AF. 1992. Stoichiometric variations of

N, P, C and O2 during a Gymnodinium catenatum red tide and their

interpretation. Mar. Ecol. Prog. Ser. 87, 123–134.

Fraga F, Ríos AF, Perez FF, Figueiras F. 1998. Theoretical limits of oxygen:

carbon and oxygen: nitrogen ratios during photosynthesis and

mineralisation of organic matter in the sea. Sci. Mar. 62, 161–168.

Fraga F. 2001. Phytoplankton biomass synthesis: application to deviations

from Redfield stoichiometry. Sci. Mar. 65, 153–169.

Fraga S, Bakun A. 1993. Global climate change and harmful algae blooms:

the example of Gymnodinium catenatum on the Galician coast, in:

Toxic phytoplankton blooms in the sea, Smayda TJ, Shimizu Y (Eds),

Elsevier, pp. 59–65.

Frouin R, Fiúza AFG, Ambar I, Boyd TJ. 1990. Observations of a poleward

current off the coasts of Portugal and Spain during winter. J.

Geophys. Res. 95(C1), 679–691.

Page 321: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

303

Gago J, Álvarez–Salgado XA, Gilcoto M, Pérez FF. 2003. Assessing the fate

of dissolved and suspended organic carbon in a coastal upwelling

system (Ría de Vigo, NW Iberian Peninsula). Estuar. Coast. Shelf Sci.

56, 271–279.

Gago J , Álvarez-Salgado XA , Nieto-Cid M , Brea S , Piedracoba S. 2005.

Continental inputs of C, N, P and Si species to the Ría de Vigo (NW

Spain). Estuar. Coast. Shelf Sci. 65, 74-82.

Gallegos CL. 1989. Microzooplankton grazing on phytoplankton in Rhode

River, Maryland: nonlinear feeding kinetics. Mar. Ecol. Prog. Ser. 57,

23-33.

García–Górriz E, Font J, Candela J. 1997. Data quality control for Vessel

Mounted Acoustic Doppler Current Profiler. Application for the

Western Mediterranean Sea. Sci. Mar. 61, 417–430.

Garvine RW. 1985. A simple model of estuarine subtidal fluctuations

forced by local and remote wind stress. J. Geophys. Res. 90(C6),

11945–11948.

Gilcoto M, Álvarez–Salgado XA, Pérez FF. 2001. Computing optimum

estuarine residual fluxes with a multiparameter inverse model

method (OERFIM): application to the Ría de Vigo (NW Spain). J.

Geophys. Res. 106, 31303–31318.

Gilcoto M. 2004. El campo termohalino como trazador de la dinámica de

la Ría de Vigo y su acoplamiento meteorológico, Ph.D. Thesis, p.,

Univ. of Vigo, 211pp.

Gilcoto M, Pardo PC, Álvarez-Salgado XA, Pérez FF. Exchange Fluxes

Between the Ría de Vigo and the Shelf: a Bidirectional Flow Forced

by Remote Wind. J. Geophys. Res., in press.

Page 322: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

304

Gill AE. 1982. Atmosphere–Ocean Dynamics. International Geophysics

Series, vol 30, edited by W. L. Donn, Academic Press, INC., London.

Godin G. 1972. The Analysis of Tides. Liverpool Univ. Press, Liverpool,

UK.

Gómez Gallego J. 1971. Estudio de las condiciones oceanográficas en la

Ría de Arosa en verano. Bol. Inst. Esp. Oceanogr. 147, 3–39.

Gómez-Gesteira M, Montero P, Prego R, Taboada JJ, Leitao P, Ruíz-

Villareal M, Neves R, Pérez-Villar C. 1999. A 2-dimensional particle

tracking model for pollution dispersion in A Coruña and Vigo Rias

(NW Spain). Oceanol. Acta 22, 167-177.

González N, González JJ, García C, Cabanas JM. 1979. Dinámica de

nutrientes en las rías de Arosa y Muros (NO de España). Bol. Inst.

Esp. Oceanogr. 5, 53–79.

Griffiths G. 1994. Using 3DF GPS Heading for Improving Underway

ADCP Data. J. Atm. Oce. Tech. 11, 1135–1143.

Hansell DA, Ducklow HW, Macdonald AM, Baringer MON. 2004.

Metabolic poise in the North Atlantic Ocean diagnosed from organic

matter transports. Limnol. Oceanogr. 49(4), 1084–1094.

Harrison WG. 1980. Nutrient regeneration and primary production in the

sea. In: Primary productivity in the sea. Falkowski PG (Ed). Plenum

Press, New York, 433–460.

Haynes R, Barton ED. 1990. A poleward flow along the Atlantic coast of

the Iberian Peninsula. J. Geophys. Res. 95, 11425–11441.

Haynes R, Barton ED. 1991. Lagrangian observations in the Iberian

Coastal Transition Zone, J. Geophys. Res. 96, 14731–14741.

Page 323: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

305

Haynes R, Barton ED, Pilling I. 1993. Development, Persistence, and

variability of upwelling filaments off the Atlantic coast of the Iberian

Peninsula. J. Geophys. Res. 98, C12: 22681–22692.

Herrera JL, Piedracoba S, Varela RA, Rosón G. 2005. Spatial analysis of the

wind field on the western coast of Galicia from accurate

measurements. Cont. Shelf Res. 25, 1728–1748.

Hidy GM. 1972. A view of recent air–sea interaction research. Bull. Am.

Meteorol. Soc. 53, 1083– 1102.

Hill AE, Hickey BM, Shillington FA, Strub PT, Brink KH, Barton ED,

Thomas AC. 1998. Eastern ocean boundaries coastal segment. In:

The Sea, vol 11. Robinson AR, Brink KH. John Wiley & Sons, New

York, pp. 29–67.

Hunter JR, Hearn CJ. 1987. Lateral and vertical variations in the wind–

driven circulation in long, shallow lakes. J. Geophys. Res. 92(C12),

13106–13114.

Huzzey L, Brubaker JM. 1988. The formation of longitudinal fronts in a

coastal plain estuary. J. Geophys. Res. 93, 1329–1334.

Ianniello J, 1979. Tidally induced residual currents in estuaries of variable

breadth and depth. J. Phys. Oceanogr. 9, 962–974.

Janzen CD, Wong K–C. 2002. Wind forced dynamics at the estuary–shelf

interface of a large coastal plain estuary. J. Geophys. Res. 107(C10),

3138, doi:10.1029/2001JC000959.

Jeffrey SW. 1997. Chlorophyll and carotenoid extinction coefficients.In:

Phytoplankton pigments in oceanography. Jeffrey SW, Mantoura

RFC, Wright SW (Eds). UNESCO publishing, Paris, pp. 595–596.

Page 324: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

306

Jeffrey SW, Wright SW. 2006. Photosynthetic pigments in marine

microalgae: insights from cultures and the sea. In: Algal cultures,

analogues of blooms and applications (1) ch 2. Subba Rao DV (Ed).

Science Publishers, Enfield (NH) USA, pp. 33-90.

Johnson TO, Smith WO. 1986. Sinking rates of phytoplankton assemblages

in the Weddell Sea marginal ice zone. Mar. Ecol. Prog. Ser. 33, 131–

137.

Joint I, Inall M, Torres R, Figueiras FG, Álvarez-Salgado XA, Rees AP,

Woodward. 2001. Two Lagrangian experiments in the Iberian

upwelling system: tracking an upwelling event and an off-shore

filament. Prog. Oceanogr. 51, 221–248.

Joint I, Groom SB, Wollast R, Chou L, Tilstone GH, Figueiras FG, Loijens

M, Smyth TJ. 2002. The response of phytoplankton production to

periodic upwelling and relaxation events at the Iberian shelf break:

estimates by the 14C method and by satellite remote sensing. J. Mar.

Sys. 32, 219–238.

Joyce TM. 1989. On in situ calibration of shipboard ADCPs. J. Atm. Oce.

Tech. 6, 169–172.

Kjerfve B. 1978. Bathymetry as an indicator of net circulation in well–

mixed estuaries. Limnol. Ocenogr. 23, 311–323.

Kjerfve B, Proehl J. 1979. Velocity variability in a cross–section of a well–

mixed estuary. J. Mar. Res. 37, 409–418.

Kjerfve B, Stevenson LH, Powell JA, Crzanowski TH, Kitchens WM. 1981.

Estimation of material fluxes in an estuarine cross section. A critical

analysis of spatial measurement density and errors. Limnol.

Ocenogr. 26, 325–335.

Page 325: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

307

Kelly KA, Lagerloef GSE, Bernstein RL. 1988. Comment on “Empirical

orthogonal function analysis of Advanced very High Resolution

Radiometer surface temperature patterns in the Santa Barbara

channel. J. Geophys. Res. 93 (C12), 15752–15754.

Kirchman DL, Ducklow HW, McCarthy JJ, Garside C. 1994. Biomass and

nitrogen uptake by heterotrophic bacteria during the spring

phytoplankton bloom in the North Atlantic Ocean. Deep-Sea Res.

41, 879-895

Klinck J, O’Brien J, Svendsen H. 1981. A simple model of fjord and coastal

circulation interaction. J. Phys. Oceanogr. 11, 1612–1626.

Knauer GA, Martin JH, Bruland KW. 1979. Fluxes of particulate carbon,

nitrogen and phosphorous in the upper water column of the

northeast Pacific. Deep-Sea Res. 26, 97-108.

Krauss W. 1986. The North Atlantic Current. J. Geophys. Res. 94, 6159–

6168.

Kundu PK, Allen JS. 1976. Some three–dimensional characteristics of low

frequency current fluctuations near the Oregon coast. J. Phys.

Oceanogr. 6, 181–199.

Laevastu T. 1963. Energy exchange in the North Pacific, its relations to

weather and its oceanographic consequences. Part 1: Formulas and

monograms for computation of heat exchange components over the

sea. Hawaiian Institute for Geophysical Research. Report 29:15.

Landry MR, Hasset RP. 1982. Estimating the grazing impact of marine

micro–zooplankton. Mar. Biol. 67, 283–288.

Page 326: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

308

Landry MR, Brown SL, Campbell L, Constantinou J, Liu H. 1998. Spatial

patterns in phytoplankton growth and microzooplankton grazing in

the Arabian Sea during monsoonal forcing. Deep–Sea Res. II 45,

2353–2368.

Laws EA. 1991. Photosynthetic quotients, new production and net

community production in the open ocean. Deep Sea Res. 38, 143–

167.

Legendre L, Gosselin M. 1989. New production and export of organic

matter to the deep ocean: consequences of some recent discoveries.

Limnol. Oceanogr. 34, 1374-1380.

Legendre L, Rassoulzadegan F. 1995. Plankton and nutrient dynamics in

marine waters. Ophelia 41, 153-172.

Longhurst AR. 1991. Role of the Marine Biosphere in the Global Carbon

Cycle. Limnol. Oceanogr. 36, 1507-1526.

López–Jamar E, Cal RM, González G, Hanson RB, Rey J, Santiago G,

Tenore KR. 1992. Upwelling and outwelling effects on the benthic

regime of the continental shelf off Galicia, NW Spain. J. Mar. Res. 50,

465–488.

Lorenzo LM, Arbones B, Tilstone GH, Figueiras FG. 2005. Across–shelf

variability of phytoplankton composition, photosynthetic

parameters and primary production in the NW Iberian upwelling

system. J. Mar. Syst. 54, 157–173.

Luek RG, Youyu L. 1997. The logarithmic layer in a tidal channel. Cont.

Shelf Res. 17, 1785–1801.

Malone TC, Kemp WM, Ducklow HW, Boynton WR, Tuttle JH, Jonas RB.

1986. Lateral variation in the production and fate of phytoplankton

in a partially stratified estuary. Mar. Ecol. Prog. Ser. 32, 149–160.

Page 327: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

309

Marañón E, Cermeño P, Fernández E, Rodríguez J, Zabala L. 2004.

Significance and mechanisms of photosynthetic production of

dissolved organic carbon in a coastal eutrophic ecosystem. Limnol.

Oceanogr. 49, 1652–1666.

Margalef R. 1978. Phytoplankton communities in upwelling areas. The

example of NW Africa. Oecol. Aquat. 97–132.

Martins CS, Hamann M, Fiúza AFG. 2002. Surface circulation in the

eastern North Atlantic, from drifters and altimetry. J. Geophys. Res.

107, C12, doi: 10.1029/2000JC000345.

Mazé JP, Arhan M, Mercier H. 1997. Volume budget of the eastern

boundary layer off the Iberian Peninsula. Deep–Sea Res. II 44, 1543–

1574.

McCreary JP, Shetye SR, Kundu PJ. 1986. Thermohaline forcing of eastern

boundary currents An application to the circulation off the west

coast of Australia. J. Mar. Res. 44, 71–92.

Middelburg JJ, Vlug T, Jaco F, van der Nat WA. 1993. Organic matter

mineralization in marine systems. Global Planet. Change 8, 47–58.

Míguez BM, Fariña-Busto L, Figueiras FG, Perez FF. 2001a. Succession of

phytoplankton assemblages in relation to estuarine hydrodynamics

in the Ría de Vigo: A box model approach. Sci Mar. 65, 65-76.

Míguez BM, Pérez FF, Souto C, Fariña–Busto L. 2001b. Flujos residuales

de intercambio entre la Ría de Vigo y la plataforma continental. Fís.

Tierra 13, 119–137.

Míguez BM. 2003. Descripción dinámica de la circulación en dos Rías

Baixas: Vigo y Pontevedra. Ph.D. Thesis, Univ. of Vigo, 182pp.

Millard RC, Owens WB, Fofonoff NP. 1990. On the calculations of the

Brunt–Väisälä frequency. Deep–Sea Res. 37, 167–181.

Page 328: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

310

Millero FJ, Sohn ML. 1992. Chemical Oceanography. CRC Press, Boca

Raton, Fla.

Myklestad SM, Skanoy E, Hestmann S. 1997. A sensitive and rapid

method for analysis of dissolved mono– and polysaccharides in

seawater. Mar. Chem. 56, 279–286.

Moncoiffé G, Álvarez–Salgado XA, Figueiras FG, Savidge G. 2000.

Seasonal and short–time–scale dynamics of microplankton

community production and respiration in an inshore upwelling

system. Mar. Ecol. Prog. Ser. 196, 111–126.

Montero P, Gómez–Gesteira M, Taboada JJ, Ruíz–Villareal M, Santos AP,

Neves RR, Prego R, Pérez–Villar V. 1999. On residual circulation of

the Ría of Vigo using a 3D baroclinic model. Bol. Inst. Esp.

Oceanogr. 15 (1–4), 31–38.

Munk WH, Anderson ER. 1948. Notes on the theory of the thermocline. J.

Mar. Res. 3, 276-295.

Neuer S, Cowles TJ. 1994. Protist herbivory in the Oregon upwelling

system. Mar. Ecol. Prog. Ser. 113, 147–162.

Nieto–Cid M, Álvarez–Salgado XA, Brea S, Pérez FF. 2004. Dissolved and

particulate carbohydrates cycling in a coastal upwelling system

(NW Iberian Peninsula). Mar. Ecol. Prog. Ser. 283, 39-54.

Nieto–Cid M, Álvarez–Salgado XA, Gago J, Pérez FF. 2005. DOM

fluorescence, a tracer for biogeochemical processes in a coastal

upwelling system (NW Iberian Peninsula). Mar. Ecol. Prog. Ser. 297,

33–50.

Nieto–Cid M, Álvarez–Salgado XA, Pérez FF. 2006. Microbial and

photochemical reactivity of fluorescent disolved orgamic matter in a

costal upwelling system. Limnol. Oceanogr. 51, 1391–1400.

Page 329: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

311

Nogueira E, Pérez FF, Ríos AF. 1997. Seasonal and long–term trends in an

estuarine upwelling ecosystem (Ría de Vigo, NW Spain). Estuar.

Coast. Shelf Sci. 44, 285–300.

Nogueira E, Ibanez F, Figueiras FG. 2000. Effect of meteorological and

hydrographic disturbances on the microplankton community

structure in the Ria de Vigo (NW Spain). Mar. Ecol. Prog. Ser. 203,

23–45.

Normann B, Zweifel UL, Hopkinson CS, Fry B. 1995. Production and

utilization of dissolved organic carbon during an experimental

diatom bloom. Limnol. Oceanogr. 40, 898–907.

Officer CB. 1980. Box models revisited. In: Estuarine and Wetlands

Processes with Emphasis on Modelling. Hamilton P, McDonald KB

(Eds), Plenum, New York, pp 65–114.

Olli K, Riser CW, Wassmann P, Ratkova T, Arashkevich E, Pasternak A.

2001. Vertical flux of biogenic matter during a Lagrangian study off

the NW Spanish continental margin. Prog. Oceanogr. 51, 443-466.

Otto L.1975. Oceanography of the Ría de Arousa (NW Spain), Rep.96, , R.

Neth. Meteorol. Inst. (KNMI), De Bilt, Netherlands, 210 pp.

Pardo PC, Gilcoto M, Pérez FF. 2001. Short–time scale coupling between

thermohaline and meteorological forcing in the Ría de Pontevedra.

Sci. Mar. 65, (Suppl.1), 229–240.

Pascual JR, Calpena V. 1985. La corriente de marea en la Ría de Arosa

(Galicia, España). Revista de Geofísica 41, 149–158.

Pascual JR. 1986. Un modelo numérico de circulación horizontal

producida por el viento y la marea en la Ría de Arosa (Galicia; NW

de España). Ph.D. Thesis, Univ. of Madrid, 156 pp.

Page 330: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

312

Pazos Y, Figueiras FG, Álvarez–Salgado XA, Rosón G. 1995. The control of

succession in red tide species in the Ría de Arousa (NW Spain) by

upwelling and stability. Harmful Marine Algal Blooms, 645–650.

Peliz A, Rosa T, Santos AMP, Pissarra JL. 2002. Jets, eddies and

counterflows in the Western Iberian upwelling system J. Mar. Sys.

35(1-2),61-77.

Peliz A, Dubert J, Santos AMP, Oliveira PB, Le Cann B. 2005. Winter

upper ocean circulation in the Western Iberian Basin–Fronts, eddies

and poleward flows: an overview. Deep Sea Res. I 52, 621–646.

Pérez FF, Fraga F. 1987. A precise and rapid analytical procedure for

alkalinity determinations. Mar. Chem. 21, 169–182.

Pérez FF, Ríos AF, King BA, Pollard RT. 1995. Decadal changes of the 0–S

relationship of the Eastern North Atlantic Central Water. Deep–Sea

Res. I 42, 1849–1864.

Pérez FF, Pollard RT, Read JF, Valencia V, Cabanas JM, Ríos AF. 2000a.

Climatological coupling of the thermohaline decadal changes in

Central Water of the Eastern North Atlantic. Sci. Mar. 64, 347–353.

Pérez FF, Álvarez–Salgado XA, Rosón G. 2000b. Stochiometry of nutrients

(C, N, P and Si) consumption and organic matter production in a

coastal inlet affected by upwelling. Mar. Chem. 69, 217–236.

Pérez FF, Castro CG, Álvarez–Salgado XA, Ríos AF. 2001. Coupling

between the Iberian basin–scale circulation and the Portugal

boundary current system. A chemical study. Deep-Sea Res. I 48,

1519–1533.

Page 331: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

313

Piedracoba S, Álvarez–Salgado XA, Rosón G, Herrera JL. 2005. Short–

timescale thermohaline variability and residual circulation in the

central segment of the coastal upwelling system of the Ría de Vigo

(northwest Spain) during four contrasting periods. J. Geophys. Res.

110, doi:10.1029/2004JC002556. Chapter 2 of this thesis.

Piedracoba S, Souto C, Gilcoto M, Pardo PC. 2005b. Hydrography and

dynamics of the Ría de Ribadeo (NW Spain), a wave driven estuary.

Estuar. Coast. Shelf Sci. 65, 726-738.

Piedracoba S, Herrera JL, Rosón G, Gilcoto M, Souto C, Álvarez–Salgado

XA. Lateral variability in the coastal upwelling system of the Ría de

Vigo (NW Spain), submitted to J. Mar. Sys. Chapter 3 of this thesis.

Piedracoba S, Nieto-Cid M, Souto , Gilcoto M, Rosón G, Álvarez-Salgado

XA, Varela R, Figueiras FG. Physical-biological coupling in the

coastal upwelling system of the Ría de Vigo (NW Spain). I: An in

situ approach, submitted to Mar. Ecol. Prog. Ser. Chapter 5 of this

thesis.

Piedracoba S, Nieto-Cid M, Teixeira IG, Garrido JL, Álvarez-Salgado XA,

Rosón G, Castro CG, Pérez FF. Physical-biological coupling in the

coastal upwelling system of the Ría de Vigo (NW Spain). I: An in

vitro approach, submitted to Mar. Ecol. Prog. Ser. Chapter 6 of this

thesis.

Pierce SD, Joyce TM. 1988. Gulf Stream velocity structure through

inversion of hydrographic and acoustic Doppler data. J. Geophys.

Res. 93(C3), 2227–2236.

Pingree RD. 1994. Winter warming in the southern Bay of Biscay and

lagrangian eddy kinematics from a deep–drogued Argos buoy. J.

Mar. Biol. Assoc. U.K. 74, 107–128.

Page 332: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

314

Pingree RD, Le Cann B. 1989. Celtic and Armorican slope and shelf

residual currents. Prog. Oceanogr. 23, 303–338.

Pingree RD, Le Cann B. 1990. Structure, strength and seasonality of the

slope currentsin the Bay of Biscay region. J. Mar. Biol. Assoc. UK 70,

857–885.

Pitcher GC, Walker DR, Mitchell–Innes BA. 1989. Phytoplankton sinking

rate dynamics in the southern Benguela upwelling system. Mar.

Ecol. Prog. Ser. 55, 261–269.

Pohlmann T. 1996. Calculating the annual cycle of the vertical eddy

viscosity in the North Sea with a three–dimensional baroclinic shelf

sea circulation model. Cont. Shelf Res. 16, 147–161.

Pollard RT, Pu S. 1985. Structure and ventillation of the upper Atlantic

Ocean northeast of the Azores. Prog. Oceanogr. 14, 443–462.

Pollard R, Read J. 1989. A Method for Calibrating Ship mounted Acoustic

Doppler Profiles and the Limitations of the Gyro Compasses. J. Atm.

Oce. Tech. 6, 859–865.

Porter KG, Feig YS. 1980. The use of DAPI for identifying and counting

aquatic microflora. Limnol. Oceanogr. 25, 943-948.

Prego R, Fraga F. 1992. A simple model to calculate the residual flows in a

Spanish Ria. Hydrographic consequences in the Ria of Vigo. Estuar.

Coast. Shelf Sci. 34, 603–615.

Prego R. 1993a. General aspects of carbon biogeochemistry in the Ría of

Vigo, northwestern Spain. Geochim. Cosmochim. Acta 57, 2041–

2052.

Prego R. 1993b. Biogeochemical pathways of phosphate in Galician Ría

(North/Western Iberian Peninsula). Estuar. Coast. Shelf Sci. 37, 437–

451.

Page 333: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

315

Prego R. 1994. Nitrogen interchanges generated by biogeochemical

processes in a Galician Ria. Mar. Chem. 45, 167–176.

Prego R, Bao R, Howland R. 1995. The biochemical cycling of dissolved

silicate in a Galician Ría. Ophelia 42, 301–318.

Prego R, Bao R. 1997. Upwelling influence on the Galician coast: silicate in

shelf water and underlying surface sediments. Cont. Shelf. Res. 17,

307–318.

Prego R. 2002. Nitrogen fluxes and budget seasonality in the Ria of Vigo

(NW Iberian Peninsula). Hydrobiologia 475/476, 161–171.

Prego R, Fraga F. 1992. A simple model to calculate the residual flows in a

Spanish ría. Hydrographic consequences in the Ría of Vigo. Estuar.

Coast. Shelf Sci. 34, 603–615.

Pritchard, D. W. 1952. Salinity distribution and circulation in the

Chesapeake Bay estuarine system. J. Mar. Res. 11, 106–123.

Pritchard DW. 1956. The dynamic structure of a coastal plain estuary. J.

Mar. Res. 15, 33–42.

Pugh DT. 1996. Tides, surges and mean-sea level. John Wiley & Sons Ltd.,

Chichester, 472pp.

Putt M, Stoecker DK. 1989. An experimental determined carbon: volume

ratio for marine ‘oligotrichous’ ciliates from estuarine and coastal

waters. Limnol. Oceanogr. 34, 1097-1103.

Quiñones RA, Platt T. 1991. The relationship between the f-ratio and the

P:R ratio in the pelagic ecosystem. Limnol. Oceanogr. 36, 211-213.

Page 334: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

316

Reed RE, Glasgow HB, Burkholder JM, Brownie C. 2004. Seasonal

physical–chemical structure, and acoustic Doppler current profiler

flow patterns over multiple years in a shallow, stratified estuary,

with implications for lateral variability. Estuar. Coast. Shelf Sci. 60,

549–566.

Reynolds–Fleming JV, Luettich RA. 2004. Wind–driven lateral variability

in a partially mixed estuary. Estuar. Coast. Shelf Sci. 60, 395–407.

Richardson TL, Cullen JJ. 1995. Changes in buoyancy and chemical

composition during growth of a coastal marine diatom: ecological

and biogeochemical consequences. Mar. Ecol. Prog. Ser. 128, 77–99.

Riebesell U. 1989. Comparison of sinking and sedimentation rate

measurements in a diatom winter /spring bloom. Mar. Ecol. Prog.

Ser. 54, 109–119.

Ríos AF. 1992. El fitoplancton de la Ría de Vigo y sus condiciones

ambientales, Ph.D. Thesis, Univ. of Santiago, 416pp.

Ríos AF, Pérez FF, Fraga F. 1992a. Water masses in upper and middle

North Atlantic Ocean east of Azores. Deep Sea Res. 39, 645–658.

Ríos AF, Nombela M, Pérez FF, Rosón G, Fraga F. 1992b. Calculation of

runoff to an estuary. Ría de Vigo. Sci. Mar. 56, 29–33.

Ríos AF, Fraga F, Pérez FF, Figueiras FG. 1998. Chemical composition of

phytoplankton and particulate organic matter in Ría de Vigo (NW

Spain). Sci. Mar. 62, 257–271.

Rosón G, Álvarez–Salgado XA, Pérez FF. 1997. A non stationary box

model to determine residual fluxes in a partially mixed estuary,

based on both thermohaline properties. Application to the Ria de

Arousa (NW Spain). Estuar. Coast. Shelf Sci. 44, 249–262.

Page 335: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

317

Rosón G, Álvarez–Salgado XA, Pérez FF. 1999. Carbon cycling in a large

coastal embayment, affected by wind–driven upwelling: short–

time–scale variability and spatial differences. Mar. Ecol. Prog. Ser.

176, 215–230.

Rodríguez F, Pazos Y, Maneiro J, Zapata M. 2003. Temporal variation in

phytoplankton assemblages and pigment composition at a fixed

station of the Ría of Pontevedra (NW Spain). Estuar. Coast. Shelf Sci.

58, 499–515.

Schroeder WW, Wiseman WJ. 1986. Low frequency shelf–estuarine

exchange processes in Mobile Bay and other estuarine systems on

the northern Gulf of Mexico”. In: Estuarine Variability, Wolfe DA

(Ed), Academic Press, San Diego, CA, pp. 355–367.

Serret P, Robinson C, Fernandez E, Teira E, Tilstone G. 2001. Latitudinal

variation of the balance between plankton photosynthesis and

respiration in the E Atlantic Ocean. Limnol. Oceanogr. 46, 1642-1652.

Sharp JH, Beauregard AY, Burdige D, Cauwet G, Curless SE, Lauck R,

Nagel K, Ogawa H, Parker AE, Primm O, Pujo–Pay M, Savidge WB,

Seitzinger S, Spyres G, Styles R. 2004. A direct instrument

comparison for measurement of total dissolved nitrogen in

seawater. Mar. Chem. 84, 181–193.

Sherr E, Sherr B. 1988. Role of microbes in pelagic food webs: A revised

concept. Limnol. Oceanogr. 33, 1225–1227.

Signell RP, Beardsley RC, Graber HC, Capotondi A. 1990. Effect of wave–

current interaction on wind–driven circulation in narrow, shallow

embayments. J. Geophys. Res. 95 (C6), 9671–9678.

Page 336: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

318

Smayda TJ, Boylen BJ. 1966. Experimental observations on the flotation of

marine diatoms. II. Skeletonema costatum and Rhizosolenia setigera.

Limnol. Oceanogr. 11, 18–34.

Smayda TJ. 1970. The suspension and sinking of phytoplankton in the sea.

Oceanogr. Mar. Biol. A. Rev. 8, 353–414.

Smith SV, Hollibaugh JT. 1993. Coastal metabolism and the oceanic

organic carbon balance. Rev. Geophys. 31, 75-89.

Smith SV, Hollibaugh JT. 1997. Annual cycle and interannual variability of

ecosystem metabolism in a temperate climate embayment. Ecol.

Monographs 67, 509–533.

Sokal RR, Rohlf FJ. 1994. Biometry: The Principles and Practice of Statistics

in Biological Research, Hardcover Ed.

Souto C. 2000. Predicción numérica y contraste experimental de la

circulación en la Ría de Vigo, Ph.D. Thesis. Univ. of Vigo, 178 pp.

Souto C, Fariña-Busto L, Álvarez E, Rodríguez I. 2001. Wind and tide

current prediction using a 3D finite difference model in the Ría de

Vigo (NW Spain). Sci. Mar. 65, 269–276.

Souto C, Gilcoto M, Fariña–Busto L, Pérez FF. 2003. Modeling the residual

circulation of a coastal embayment affected by wind–driven

upwelling: Circulation of the Ría de Vigo (NW Spain). J. Geophys.

Res. 108(C11), 3340, doi: 10.1029/2002JC001512.

Spencer CP. 1975. The micronutrients elements. In: Chemical

oceanography, 2nd edn, Vol 2, Riley JP, Skirrow G (Eds). Academic

Press, London, pp. 245-300.

Spencer CP. 1983. Marine biogeochemistry of Silicon. In: Silicon

geochemistry and biogeochemistry. Aston SR (Ed). Academic Press,

San Diego, 101–139.

Page 337: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

319

Strathmann R. 1967. Estimating the organic carbon content of

phytoplankton from cell volume or plasma volume. Limnol.

Oceanogr. 12, 411-418.

Strickland JD, Parsons TR. 1972. A practical handbook of seawater

analysis. 2nd ed. Bull. Fish. Res. Bd. Can, 167 pp.

Svendsen H, Thomson R. 1978. Wind driven circulation in a fjord. J. Phys.

Oceanogr. 8, 703–713.

Sy A. 1988. Investigation of large scale circulation patterns in the central

North Atlantic: the North Atlantic Current, the Azores Current, and

the Mediterranean Water Plume in the area of the Mid–Atlantic

Ridge. Deep Sea Res. 35, 383–413.

Teira E, Serret P, Fernández E. 2001a. Phytoplankton size–structure,

particulate and dissolved organic carbon production and oxygen

fluxes through microbial communities in the NW Iberian coastal

transition zone. Mar. Ecol. Prog. Ser. 219, 65–83.

Teira E, Pazó MJ, Serret P, Fernández E. 2001. Dissolved organic carbon

production by microbial populations in the Atlantic Ocean. Limnol.

Oceanogr. 46, 1370–1377.

Teira E, Abalde J, Álvarez–Osorio MT, Bode A, Cariño C, Cid A,

Fernández E, González N, Lorenzo J, Valencia J, Varela M. 2003.

Plankton carbon budget in a coastal wind–driven upwelling station

off A Coruña (NW Iberian Peninsula). Mar. Ecol. Prog. Ser. 265, 31–

43.

Page 338: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

320

Tenore KR, Bouer LF, Cal RM, García Fernández C, González N,

González–Gurriaran E, Hanson RB, Iglesias J, From M, López–Jamar

E, McClain J, Pamatmat MM, Pérez A, Rhoads DC, Santiago GD,

Tietjen J, Westrich J, Windom HL. 1982. Coastal upwelling in the

Rías Bajas, NW Spain: contrasting the benthic regimes in the Rías of

Arosa and Muros. J. Mar. Res. 40, 701–772.

Tett P, Edwards A. 1984. Mixing and plankton : an interdisciplinary theme

in oceanography. Oceanog. Mar. Bio. Annual Rev. 22, 99-123.

Thronton DCO, Thake B. 1998. Effect of temperature on the aggregation of

Skeletonema costatum (Bacillariophyceae) and the implication fro

carbon flux in coastal waters. Mar. Ecol. Prog. Ser. 174, 223–231.

Tilstone GH, Figueiras FG, Fermín EG, Arbones B. 1999. The significance

of nano–phytoplankton photosynthesis and primary productivity in

a coastal upwelling ecosystem (Ría de Vigo, NW Spain). Mar. Ecol

Prog. Ser. 205, 23–41.

Tilstone GH, Míguez BM, Figueiras FG, Fermín EG. 2000. Diatom

dynamics in a coastal ecosystem affected by upwelling: coupling

between species succession, circulation and biogeochemical

processes. Mar. Ecol. Prog. Ser. 205, 23–41.

Tilstone GH, Figueiras FG, Lorenzo LM, Arbones B. 2003. Phytoplankton

composition, photosynthesis and primary production during

different hydrographic conditions at the Northwest Iberian

upwelling system. Mar. Ecol. Prog. Ser. 252, 89–104.

Torre Enciso E. 1958. Estado actual del conocimiento de las Rías Gallegas,

en Homenaxe a Otero Pedrayo, Galaxia, pp 237–250.

Page 339: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

321

Torres R, Barton ED, Miller P, Fanjul E. 2003. Spatial patterns of wind and

sea surface temperature in the Galician upwelling region. J.

Geophys. Res. 108 (C4), 27-1–27-13.

Torres–López S, Varela RA, Delhez E. 2001. Residual circulation and

thermohaline distribution of the Ría de Vigo: A 3–D hydrodynamic

model. Sci. Mar. 65, 277–289.

Torres-López S, Álvarez-Salgado XA, Varela RA. 2005. Offshore export

versus in situ fractionated mineralization: a 1-D model of the fate of

the primary production of the Rías Baixas (Galicia, NW Spain). J.

Mar. Sys. 54, 175– 193.

UNESCO. 1985. The international system of units (SI) in oceanography.

UNESCO Technical Papers in Marine Sciences 45, 1–124.

UNESCO. 1994. Protocols for the Joint Global Ocean Flux Study (JGOFS)

core measurements. Manuals and Guide, No.29. Intergovernmental

Oceanographic Commission, Paris.

Utermohl H. 1958. Zur Vervollkommung der quantitativen

Phytoplankton-Methodic. Comm. Assoc. Int. Limnol. Theor. Appl 9,

1-38.

Valle–Levinson A, Lwiza KMM. 1995. The effects of channels and shoals

on exchange between the Chesapeake Bay and the adjacent ocean. J.

Geophys. Res. 100, 18551–18563.

Valle–Levinson A. 1995. Observations of barotropic and baroclinic

exchanges in the lower Chesapeake Bay. Cont. Shelf Res. 15, 1631–

1647.

Valle–Levinson A, Li C, Royer T, Atkinson L. 1998. Flow patterns at the

Chesapeake Bay entrance. Cont. Shelf Res. 18, 1157–1177.

Page 340: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

322

Valle–Levinson A, Atkinson L. 1999. Spatial gradients in the flow over an

estuary channel. Estuaries 8, 179–193.

Valle–Levinson A, Wong KC, Lwiza KKM. 2000. Fortnightly variability in

the transverse dynamics of a coastal plain estuary. J. Geophys. Res.

105(C2), 3413–3424.

Valle–Levinson A, Blanco JL, Fierro JJ. 2002. Observations of wind effects

on Exchange flows in a channel constriction of the Chilean Inland

Sea, The 2nd Meeting on the Physical Oceanography of Sea Straits,

Villefranche, 15th–19th April.

Varela M, Fuentes E, Pernas E, Cabanas JM. 1984. Producción primaria de

las Rías Baixas de Galicia. As Rías galegas, Cuadernos da Area de

Ciencias Mariñas, Seminario de Estudos Galegos 1, 173–182.Vieira

MEC. 1985. Estimates of subtidal volume flux in mid Chesapeake

Bay. Estuar. Coast. Shelf Sci. 21, 411–427.

Varela M, Prego R, Pazos Y. 2004. Vertical biogenic particle flux in a

western Galician ria (NW Iberian Peninsula). Mar. Ecol. Prog. Ser.

269, 17–32.

Varela M, Prego R, Pazos Y, Moroño A. 2005. Influence of upwelling and

river runoff interaction on phytoplankton assemblages in a middle

Galician ria and comparison with northern and southern rias (NW

Iberian Peninsula). Estuar. Coast. Shelf Sci. 64, 721-737.

Verity P, Robertson CY et al. 1992. Relationships between cell volume and

the carbon and nitrogen content of marine photosynthetic

nanoplankton. Limnol. Oceanogr. 37, 1434-1446.

Vidal Romani JR. 1984. A orixe das rías galegas. Estado da cuestion (1886–

1983), As rias galegas, Cuadernos da Area de Ciencias Mariñas,

Seminario de Estudos Galegos 1, 13–25.

Page 341: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

323

Vieira MEC. 1986. The meteorologically driven circulation in mid–

Chesapeake Bay. J. Mar. Res. 44, 473–493.

Villarino ML, Figueiras FG, Jones KJ, Álvarez–Salgado XA, Richards J,

Edwards A. 1995. Evidence of in situ diel migration of a red–tide

microplankton species in Ría de Vigo (NW Spain). Mar. Biol. 123,

607–617.

Wada E, Hattori A. 1991. Nitrogen in the sea: forms, abundances, and rate

processes. CRC Press, Boca Raton, 208pp.

Wafar MVM, Le Corre P, L’Helguen S. 1995. f-Ratios calculated with and

without urea uptake in nitrogen uptake by phytoplankton. Deep Sea

Res. I 42, 1669-1674.

Waite A, Bienfang PK, Harrison PJ. 1992. Spring bloom sedimentation in a

subartic ecosystem. II. Succession and sedimentation. Mar. Biol. 114,

131–138.

Waldron HN, Probyn TA, Brundrit GB. 1998. Carbon pathways and

export associated with the southern Benguela upwelling system: a

re-appraisal. South African J. Mar. Sci. 19, 113-118.

Walsh JJ. 1991. Importance of continental margins in the marine

biogeochemical cycling of carbon and nitrogen. Nature 359, 53–55.

Walters RA. 1982. Low–frequency variations in sea level and currents in

South Francisco Bay. J. Phys. Oceanogr. 12, 658–668.

Walters RA, Gartner JW. 1985. Subtidal sea level and current variations in

the northern reach of San Francisco Bay. Estuar. Coast. Shelf Sci. 21,

17–32.

Wang JD, Elliott AJ. 1978. Non–tidal variability in the Chesapeake Bay in

the Potomac River: evidence for non–local forcing. J. Phys.

Oceanogr. 8, 225–232.

Page 342: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Chapter 8

324

Weisberg RH. 1976. A note on estuarine mean flow estimation. J. Mar. Res.

34 (3).

Weisberg RH, Sturges W. 1976. Velocity Observations in the West Passage

of Narragansett Bay: A partially mixed estuary. J. Phys. Oceanogr. 6,

345–354.

Weiss GM, Harding Jr. LW, Itsweire EC, Campbell JC. 1997. Assessing

lateral variations of surface chlorophyll concentrations in the

Chesapeake Bay using aircraft remote sensing. Mar. Ecol. Prog. Ser.

149, 183–199.

Wieringa J. 1986. Roughness dependent geographical interpolation of

surface winds speed averages. Quart. J. Roy. Meteor. Soc. 112, 867–

889.

Williams PJleB. 1998. The balance of plankton respiration and

photosynthesis in the open oceans. Nature 394, 55-57.

Wollast R. 1993. Interactions of carbon and nitrogen cycles in the coastal

zone. In: Interaction of C, N, P and S Biochemical Cycles and global

change. Wollast R, Mackenzie FP, Chou L (Eds), Springer–Verlag,

Berlin, pp. 195–210.

Wollast R. 1998. Evaluation and comparison of the global carbon cycle in

the coastal zone and in the open ocean. In: The Sea, vol. 10. Brink

KH, Robinson AR (Eds), John Wiley & Sons, New York, pp. 213–252.

Wong K–C, Garvine RW. 1984. Observations of Wind–Induced, Subtidal

Variability in the Delaware Estuary. J. Geophys. Res. 89(C6), 10589–

10597.

Wong KC, Wilson RE. 1984. Observations of low–frequency variability in

Great South Bay and relations to atmospheric forcing. J. Phys.

Oceanogr. 14, 1893–1900.

Page 343: digital.csic.esdigital.csic.es/bitstream/10261/109805/1/Piedracoba... · TESIS DOCTORAL presentada en el Departamento de Xeociencias Mariñas e Ordenación do Territorio, Facultade

Literatura Cited

325

Wong KC, DiLorenzo J. 1988. The response of Delaware’s inland bays to

ocean forcing. J. Geophys. Res. 93, 12525–12535.

Wong KC. 1994. On the nature of transverse variability in a coastal plain

estuary. J. Geophys. Res. 99 (C7), 14209–14222.

Wong KC, Münchow A. 1995. Buoyancy forced interaction between

estuary and inner shelf: Observation. Cont. Shelf Res. 15(1), 59–88.

Wong KC, Moses–Hall JE. 1998. The tidal and subtidal variations in the

transverse salinity and current distributions across a coastal plain

estuary. J. Mar. Res. 56, 489– 517.

Wong KC. 2002. On the spatial structure of currents across the

Chesapeake and Delaware Canal. Estuaries 4A, 519-527.

Wooster WS, Bakun A, Mclain DR. 1976. The seasonal upwelling cycle

along the eastern boundary of the North Atlantic. J. Mar. Res. 34,

131–141.

Yentsch CS, Menzel DW. 1963. A method for the determination of

phytoplankton chlorophyll and phaeophytin by fluorescence. Deep–

Sea Res. 10, 221–231.

Zapata M, Rodríguez F, Garrido JL. 2000. Separation of chlorophylls and

carotenoids from marine phytoplankton: a new HPLC method using

a reversed phase C8 column and pyridine–containing mobile

phases. Mar. Ecol. Prog. Ser. 195, 29–45.

Zimmerman RC, Kremer JM, Dugdale RC. 1987. Acceleration of nutrient

uptake by phytoplankton in a coastal upwelling ecosystem: A

modeling analysis. Limnol. Oceanogr. 32, 359–367.