Author's personal copy

Similar documents
The Coriolis force, geostrophy, Rossby waves and the westward intensification

Super-parameterization of boundary layer roll vortices in tropical cyclone models

DUE TO EXTERNAL FORCES

The Surface Currents OCEA 101

Lecture 22: Ageostrophic motion and Ekman layers

Atmospheric Waves James Cayer, Wesley Rondinelli, Kayla Schuster. Abstract

Section 6. The Surface Circulation of the Ocean. What Do You See? Think About It. Investigate. Learning Outcomes

Goal: Develop quantitative understanding of ENSO genesis, evolution, and impacts

Winds and Ocean Circulations

Salmon: Introduction to ocean waves

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

Gravity waves in stable atmospheric boundary layers

Upwelling. LO: interpret effects of upwelling on production of marine ecosystems. John K. Horne University of Washington

Geophysical Fluid Dynamics of the Earth. Jeffrey B. Weiss University of Colorado, Boulder

Lecture 8 questions and answers The Biological Pump

SURFACE CURRENTS AND TIDES

Numerical Simulations of a Train of Air Bubbles Rising Through Stagnant Water

AIRFLOW GENERATION IN A TUNNEL USING A SACCARDO VENTILATION SYSTEM AGAINST THE BUOYANCY EFFECT PRODUCED BY A FIRE

ATMS 310 Tropical Dynamics

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

Variability of surface transport in the Northern Adriatic Sea from Finite-Size Lyapunov Exponents" Maristella Berta

AOS 103. Week 4 Discussion

An experimental study of internal wave generation through evanescent regions

Lesson: Ocean Circulation

LONG WAVES OVER THE GREAT BARRIER REEF. Eric Wolanski ABSTRACT

+ R. gr T. This equation is solved by the quadratic formula, the solution, as shown in the Holton text notes given as part of the class lecture notes:

Summary of Lecture 10, 04 March 2008 Introduce the Hadley circulation and examine global weather patterns. Discuss jet stream dynamics jet streams

Chapter 6: Atmospheric Pressure, Wind, and Global Circulation

Oceans and the Global Environment: Lec 2 taking physics and chemistry outdoors. the flowing, waving ocean

Computational Analysis of Oil Spill in Shallow Water due to Wave and Tidal Motion Madhu Agrawal Durai Dakshinamoorthy

Chapter 2. Turbulence and the Planetary Boundary Layer

WORKING PAPER SKJ 1 IMPACT OF ENSO ON SURFACE TUNA HABITAT IN THE WESTERN AND CENTRAL PACIFIC OCEAN. Patrick Lehodey

from a decade of CCD temperature data

EARTH, PLANETARY, & SPACE SCIENCES 15 INTRODUCTION TO OCEANOGRAPHY. LABORATORY SESSION #6 Fall Ocean Circulation

Kathleen Dohan. Wind-Driven Surface Currents. Earth and Space Research, Seattle, WA

Lecture 13 El Niño/La Niña Ocean-Atmosphere Interaction. Idealized 3-Cell Model of Wind Patterns on a Rotating Earth. Previous Lecture!

Some basic aspects of internal waves in the ocean & (Tidally driven internal wave generation at the edge of a continental shelf)

Sea and Land Breezes METR 4433, Mesoscale Meteorology Spring 2006 (some of the material in this section came from ZMAG)

Effect of Orography on Land and Ocean Surface Temperature

Investigation of Suction Process of Scroll Compressors

Week 6-7: Wind-driven ocean circulation. Tally s book, chapter 7

Geostrophic and Tidal Currents in the South China Sea, Area III: West Philippines

Training program on Modelling: A Case study Hydro-dynamic Model of Zanzibar channel

Wednesday, September 27, 2017 Test Monday, about half-way through grading. No D2L Assessment this week, watch for one next week


Alongshore wind stress (out of the page) Kawase/Ocean 420/Winter 2006 Upwelling 1. Coastal upwelling circulation

Atmospheric Rossby Waves in Fall 2011: Analysis of Zonal Wind Speed and 500hPa Heights in the Northern and Southern Hemispheres

Mountain Forced Flows

Development of Fluid-Structure Interaction Program for the Mercury Target

Midterm Exam III November 25, 2:10

The events associated with the Great Tsunami of 26 December 2004 Sea Level Variation and Impact on Coastal Region of India

Atmosphere, Ocean and Climate Dynamics Fall 2008

Lecture 18: El Niño. Atmosphere, Ocean, Climate Dynamics EESS 146B/246B

Study of Passing Ship Effects along a Bank by Delft3D-FLOW and XBeach1

Preliminary results of SEPODYM application to albacore. in the Pacific Ocean. Patrick Lehodey

IX. Upper Ocean Circulation

LOW PRESSURE EFFUSION OF GASES revised by Igor Bolotin 03/05/12

McKnight's Physical Geography 11e

2.4. Applications of Boundary Layer Meteorology

Small- and large-scale circulation

10.6 The Dynamics of Drainage Flows Developed on a Low Angle Slope in a Large Valley Sharon Zhong 1 and C. David Whiteman 2

Introduction to Oceanography OCE 1001

Lesson 14: Simple harmonic motion, Waves (Sections )

The Ocean is a Geophysical Fluid Like the Atmosphere. The Physical Ocean. Yet Not Like the Atmosphere. ATS 760 Global Carbon Cycle The Physical Ocean

Sound scattering by hydrodynamic wakes of sea animals

VERTICAL DISTRIBUTION OF PLANKTON IN THE MIXED LAYER, TURBULENCE AND CLIMATE CHANGE

El Niño Lecture Notes

Air Pressure and Wind

Proceedings of Meetings on Acoustics

Role of the oceans in the climate system

It is advisable to refer to the publisher s version if you intend to cite from the work.

LOW PRESSURE EFFUSION OF GASES adapted by Luke Hanley and Mike Trenary

SIO 210 Final examination Wednesday, December 11, PM Sumner auditorium Name:

Conditions for Offshore Wind Energy Use

3. Observed initial growth of short waves from radar measurements in tanks (Larson and Wright, 1975). The dependence of the exponential amplification

Waves, Turbulence and Boundary Layers

Unsteady Wave-Driven Circulation Cells Relevant to Rip Currents and Coastal Engineering

SIO 210 Problem Set 3 November 4, 2011 Due Nov. 14, 2011

6.28 PREDICTION OF FOG EPISODES AT THE AIRPORT OF MADRID- BARAJAS USING DIFFERENT MODELING APPROACHES

Internal Tides and Solitary Waves in the Northern South China Sea: A Nonhydrostatic Numerical Investigation

Chapter. Air Pressure and Wind

INTERNATIONAL JOURNAL OF CIVIL AND STRUCTURAL ENGINEERING Volume 1, No 4, 2010

Chapter 9: Circulation of the Ocean

Observations and Modeling of Coupled Ocean-Atmosphere Interaction over the California Current System

PROC. ITB Eng. Science Vol. 36 B, No. 2, 2004,

The Air-Sea Interaction. Masanori Konda Kyoto University

The total precipitation (P) is determined by the average rainfall rate (R) and the duration (D),

CHAPTER 7 Ocean Circulation

Factors that determine water movement. Morphometry Structure of stratification Wind patterns

EFFECTS OF WAVE, TIDAL CURRENT AND OCEAN CURRENT COEXISTENCE ON THE WAVE AND CURRENT PREDICTIONS IN THE TSUGARU STRAIT

Lecture 24. El Nino Southern Oscillation (ENSO) Part 1

ESCI 343 Atmospheric Dynamics II Lesson 10 - Topographic Waves

Internal Waves in Straits Experiment Progress Report

THE EFFECT OF THE OCEAN EDDY ON TROPICAL CYCLONE INTENSITY

Chapter 11 Waves. Waves transport energy without transporting matter. The intensity is the average power per unit area. It is measured in W/m 2.

The impacts of explicitly simulated gravity waves on large-scale circulation in the

3 The monsoon currents in an OGCM

Currents measurements in the coast of Montevideo, Uruguay

Atmospheric Rossby Waves Fall 2012: Analysis of Northern and Southern 500hPa Height Fields and Zonal Wind Speed

Ocean Currents Unit (4 pts)

Transcription:

Theoretical Population Biology 76 (9) 258 267 Contents lists available at ScienceDirect Theoretical Population Biology journal homepage: www.elsevier.com/locate/tpb The effects of abrupt topography on plankton dynamics L. Zavala Sansón a,, A. Provenzale b a Department of Physical Oceanography, CICESE, Km 17 Carretera Tijuana-Ensenada, 2286 Ensenada, Baja California, Mexico b Institute of Atmospheric Sciences and Climate, CNR, Corso Fiume 4, I-1133 Torino, Italy a r t i c l e i n f o a b s t r a c t Article history: Received 25 November 8 Available online 6 September 9 Keywords: Plankton dynamics NPZ model Bottom topography Quasi two-dimensional flow Plankton population dynamics in the upper layer of the ocean depends on upwelling processes that bring nutrients from deeper waters. In turn, these depend on the structure of the vertical velocity field. In coastal areas and in oceanic regions characterized by the presence of strong submarine topographic features, the variable bottom topography induces significant effects on vertical velocities and upwelling/downwelling patterns. As a consequence, large plankton and fish abundances are frequently observed above seamounts, canyons and steep continental shelves. In this work, the dynamics of an NPZ (nutrient-phytoplanktonzooplankton) system is numerically studied by coupling the ecosystem model with a quasi twodimensional (2D) fluid model with topography. At variance with classical 2D approaches, this formulation allows for an explicit expression of the vertical motions produced when fluid columns are squeezed and stretched as they experience changes of depth. Thus, input or output of nutrients at the surface are associated with fluid motion over the bottom topography. We examine the dynamics of a cyclonic vortex over two basic topographies: a steep escarpment and a submarine mountain. We show that plankton abundance over the escarpment is modulated by the passing of topographic Rossby waves, generated by the vortex-topography interaction. In such configuration, advection effects driven by the flow over the escarpment are of limited relevance for the dynamics of biological fields. By contrast, we find that the flow resulting from the interaction of a vortex with a seamount is sufficiently strong and persistent to allow for a remarkable increase of nutrients, and a corresponding enhancement of phytoplankton and zooplankton concentrations. Over the seamount, advection effects associated with trapped flow perturbations around the summit play an essential role. 9 Elsevier Inc. All rights reserved. 1. Introduction Plankton dynamics at the ocean surface is a key aspect of marine life, as well as in the global carbon cycle. Planktonic populations depend on light cycles and nutrient availability in the euphotic layer of the ocean, where phytoplankton fix carbon through photosynthesis and are consumed by zooplankton and higher trophic levels (Mann and Lazier, 6; Lévy, 8). There is a wide range of additional biogeochemical, dynamical and radiative mechanisms that determine the evolution of plankton abundance. These complex interactions give rise to the development of numerous theoretical approximations that couple one or more of such mechanisms. A basic physical process is the injection and extraction of nutrients into and from the surface ocean, which play a major role in the dynamics of plankton populations. Corresponding author. E-mail addresses: lzavala@cicese.mx (L. Zavala Sansón), a.provenzale@isac.cnr.it (A. Provenzale). There might be several mechanisms introducing or extracting nutrients into/from the upper ocean surface, depending on the temporal and spatial scales of sea motion. For instance, it has been suggested that the propagation of planetary, large scale Rossby waves is associated with the introduction of nutrients to surface waters (Uz et al., 1). On smaller scales of motion, mesoscale mechanisms play a major role in the exchange between the euphotic layer and the deeper waters. Some of these processes are due to mesoscale vortices, which are able to pump up or down nutrients during intensification (McGillicuddy and Robinson, 1997); another mechanism is associated with the horizontal transport of fluid columns across density fronts, which induces cyclonic (anticyclonic) structures as they are stretched (compressed) (Lévy, 3). In addition, the submesoscale motion of oceanic filaments can be responsible for intense vertical advection of nutrients and biological material (Lévy et al., 1). In this paper we focus on the supply of deep water nutrients to the ocean surface by the influence of bottom topography, and on the consequences of this process for plankton ecosystems. The scientific and economic relevance of this topic is evident when considering that large plankton and fish abundances are frequently 4-589/$ see front matter 9 Elsevier Inc. All rights reserved. doi:1.116/j.tpb.9.8.4

L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 259 observed above topographic features, as widely reviewed by Genin (4). A number of biological and physical mechanisms has been proposed to explain these observations, which are mainly related to the interaction of oceanic currents with abrupt topographic features, as well as the behavioral response of plankton to vertical currents. The main physical mechanism considered here is the upwelling of deep nutrient-rich water associated with stretching and squeezing effects as an oceanic flow impinges the topography (see e.g. Huppert and Bryan, 1976). This effect is determined by potential vorticity conservation, and, ultimately, by the rotation of the Earth. This mechanism has been invoked by a number of researchers to explain aggregations of plankton and fish on a wide variety of topographies (see e.g. Genin and Boehlert, 1985; Beckman and Mohn, 2), although some other studies have questioned its role to explain enhanced primary production, especially over seamounts (Comeau et al., 1995; Genin, 4). Here we investigate the influence of bottom topography on the temporal and spatial dynamics of biological fields at the ocean surface, by coupling simple, yet highly nonlinear, physical and biological models. Such relatively simplified models are often a useful tool for gaining a better understanding of complex processes in the ocean, and therefore this approach is adopted throughout the paper. More specifically, these formulations are used to examine the interaction of an oceanic flow with a given topographic feature (a steep slope or a seamount), and examine the consequences on plankton dynamics. The link between the physical and biological components is the horizontal divergence of the flow resulting from its interaction with the topography. Among the wide variety of topographic features, we focus our study on a mesoscale escarpment and a seamount. Steep escarpments are frequently situated at the edge of continental shelves, where aggregations of fish and cetaceans have been registered (presumably due to plankton abundance, e.g. Moulins et al., 8). Furthermore, along these locations topographic Rossby waves are easily generated by the influence of mesoscale vortices or currents (Oey and Lee, 2). Our aim is to explore the potential of these oceanic oscillations to affect plankton ecosystems at the ocean surface. Afterwards, we analyze the corresponding role of a seamount, since several studies have documented the remarkable abundance of biological species on and around these features (Genin and Boehlert, 1985; Beckman and Mohn, 2; Trasviña-Castro et al., 3). We first consider the case of the steep escarpment, where it is easier to understand the effects of topography on nutrient and plankton dynamics, and then we proceed to the case of the seamount, where the biological and physical fields show a more complicated behavior. In recent studies, the basic properties of plankton ecosystems have often been simulated by means of simple nutrientphytoplankton-zooplankton (NPZ) models coupled with a hydrodynamical model in a single, surface oceanic layer (e.g. Martin et al., 2; Pasquero et al., 5). In some cases, horizontal advection has been simulated by considering two-dimensional models, which lack an explicit expression for vertical velocity. When using these approaches, input and output of nutrients are parameterized by prescribing a vertical flux from a nutrient-rich lower layer. In the present study, the use of a shallow water model with variable topography (see Grimshaw et al., 1994) allows for an explicit non-zero horizontal divergence that can be related to vertical displacements, and then used as a source/sink term of nutrients in the NPZ model. The organization of the paper is as follows. The NPZ approach and the flow model are described in Section 2, together with the biological and physical parameters used in the simulations. In Sections 3 and 4, the evolution of the ecological model is studied under the influence of a cyclonic vortex over an escarpment and a submarine mountain, respectively. In Section 5 we give a summary and a discussion of the main results. 2. NPZ and hydrodynamical models 2.1. Biological model The ecological model is a standard NPZ formulation applied to the surface layer, equivalent to that used by Pasquero et al. (5). The evolution equations of nutrients (N), phytoplankton (P) and zooplankton (Z) are expressed as follows: dn dt dp dt dz dt = (S b + S n )(N N) β NP κ N + N [ + µ N (1 γ ) aɛp ] 2 Z a + ɛp + µ 2 P P + µ Z Z 2 2 Z = β NP κ N + N aɛp a + ɛp µ 2 P P (2) = γ aɛp 2 Z a + ɛp 2 µ Z Z 2. (3) The vertical nutrient supply is given by the first term on the right-hand side of the first equation, representing the nutrient input/output from/to a deeper layer with constant nutrient concentration N. The rate S b +S n has a component S b proportional to the vertical velocity of the flow, and a weak, constant positive input S n that maintains a small, non-zero level of the NPZ fields in absence of fluid motion; the subscript n indicates normal conditions. The rate S b can be positive or negative, depending on how the flow evolves over the bottom topography (hence the subscript b). This rate is a variable function of time and space, S b (t, x, y), and is obtained by coupling the physical model described in next subsection. This term is the key mechanism for up- and downwelling of nutrients in the surface layer and it is further discussed below. The second and third terms on the right-hand side of the first equation represent nutrient consumption by phytoplankton, characterized by the maximum rate β, and regeneration from dead phytoplankton and zooplankton through remineralization, an implicit parameterization of the bacterial loop. The regeneration term is weighted by µ N, µ N 1, which measures the efficiency with which organic material is regenerated into nutrients. Phytoplankton population dynamics is described by the second equation: it grows by consuming nutrients, it is grazed by zooplankton, and it is affected by the mortality rate µ P. The zooplankton population is governed by the third equation: it grows due to phytoplankton consumption and it has a quadratic mortality term with coefficient µ Z. An additional sink term of phytoplankton can be included. This term could be added in the second equation with the form S b P, where the rate S b refers to negative values of S b. It represents the loss of phytoplankton due to sinking into deep waters during downwelling events. However, additional simulations indicated that the effect of this sink term is small enough and it can be neglected. Despite the apparent simplicity of a standard NPZ formulation, its solutions often show rather complicated behavior, as shown by several authors (see e.g. Koszalka et al., 7). This assertion remains valid even when the model assumes homogeneous conditions, i.e. when it is not coupled with a physical model that solves horizontal motion. In order to gain a clear interpretation of the results, therefore, some effects that are usually relevant for plankton dynamics, such as phytoplankton response to light or the dynamics of detritus, are not included. Here, attention is mainly focused on the general response of NPZ fields to flow-topography interaction, rather than on proposing explanations of specific field observations. (1)

26 L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 2.2. Quasi 2D model with topography The dynamical model used here is a quasi two-dimensional formulation described by several authors (e.g. Grimshaw et al., 1994; Zavala Sansón and van Heijst, 2), which is an extended version of classical 2D models. In this approximation, the dynamical equation for the vertical component of relative vorticity, ω, has the form: ω t + J(q, ψ) = A 2 ω, (4) where 2 = 2 / x 2 + 2 / y 2 is the horizontal Laplacian, J = q/ x ψ/ y q/ y ψ/ x the Jacobian operator, and A is the horizontal eddy viscosity. The potential vorticity is q = ω + f (5) h where f is the Coriolis parameter at a given latitude, and h(x, y) is the total fluid depth, which in general depends on the horizontal coordinate (x, y) due to topographic variations, and it is time independent when using the rigid-lid approximation (Grimshaw et al., 1994). Relative vorticity can be written in terms of the transport function ψ as ω = 1 h 2 ψ + 1 h ψ. (6) h2 It is important to emphasize the effect of vertical motions for this model. Using (u, v) as the horizontal velocity, the continuity equation reads (hu) x + (hv) y which can be rewritten as =, (7) u x + v y = 1 Dh h Dt, (8) where D/Dt = / t + u / x + v / y is the material derivative. This expression states that the horizontal divergence is determined by changes in the height of fluid columns as they are advected over the topography. Vertical motion is associated with horizontal divergence: flow divergence is due to upwelling as a fluid column is squeezed when moving uphill, and flow convergence is due to downwelling as columns are stretched when a fluid moves downhill. Therefore, the vertical transport of nutrients in the NPZ model due to the flow is proportional to the horizontal divergence: S b (t, x, y) = u x + v y. (9) Negative values of S b might induce local nutrient depletion in some regions of the surface layer, while positive values of S b are associated with local nutrient injection. Clearly, the value of S b integrated over the whole domain is zero. In addition, at any given location and time it cannot be decided a priori what is the relative weight of S b compared to S n, since it depends on how the flow evolves over the topography. 2.3. Physical and biological parameters The simulations are based on a finite differences code that solves Eqs. (4) and (6) [originally developed by R. Verzicco and P. Orlandi and subsequently modified to include topography effects by Zavala Sansón and van Heijst ()]. Time advance in (4) is performed by using a third order Runge Kutta scheme. The relation between relative vorticity and the transport function, Eq. (6), is solved by using a multigrid method. The numerical domain is discretized with a 256 256 square grid with size x = 1 km, so the simulations represent a 256 km 256 km square area. The timestep is t = 36 s. No-slip boundary conditions are imposed Table 1 Numerical parameters. Symbol Value Coriolis parameter f 1 4 s 1 Eddy viscosity A m 2 s 1 Domain 256 256 km 2 Gridsize x 1. km Timestep t 36 s Vortex parameters Radius R 2 km Maximum vorticity ω 3.68 1 5 s 1 Initial position (x, y ) (156 km, km) at the vertical walls. However, the main results described here are located far enough from the boundaries, so that their effect is not relevant; additional simulations using free-slip conditions show equivalent results. In all simulations, the Coriolis parameter is chosen as f = 1 4 s 1, and the horizontal eddy viscosity is A = m 2 s 1. The initial condition is a Bessel monopolar vortex with horizontal length scale R = 2 km and maximum vorticity ω = 3.68 1 5 s 1, which gives a rotation period of approximately 4 days, and a Rossby number of Ro = ω /f.3. In all cases, the vortex is initially located in the south-eastern region of the domain, (x, y ) = (156 km, km), far enough from the walls, but relatively close to the topographic features. The duration of the simulations is one month (3 days), during which the flow experiences the influence of topography. Table 1 shows a summary of the numerical parameters used in all simulations. Two particular topographies, the steep escarpment and the seamount shown in Fig. 1, are adopted and described separately in Sections 3 and 4. In order to solve the NPZ model, the semi-lagrangian approach used by Pasquero et al. (5) is adopted. Here 256 2 particles are initially uniformly distributed in the whole domain. It is assumed that each particle represents a water parcel characterized by its own biological fields, which are varying in time according to the NPZ formulation. This means that derivatives at the left-hand side of the NPZ equations are actually material derivatives at the surface of the ocean, that is, they denote the time evolution of the biological fields following a fluid parcel. Therefore, advection of particles by the flow accounts for advection of NPZ concentrations. These fields are assumed to be independent of concentrations at neighboring particles, i.e. there is no mixing among nearby parcels (Pasquero et al., 5). Besides, the biological fields are passive, i.e. they do not exert any influence on the flow motion. Of course, there must be enough particles to cover the whole domain (that is, there must always exist at least 4 to 5 particles near every gridpoint) in order to interpolate the NPZ fields onto the regular grid. This condition is ensured by the relative weakness of the divergence field (or, in other words, by the smallness of vertical displacements compared with horizontal motions). The NPZ fields are not allowed to have negative values, so they are forced to be zero when necessary. The biological parameters used in the simulations are presented in Table 2 and lead to the presence of a single, globally stable fixed point of the homogeneous system. The parameter values are essentially the same as those in Pasquero et al. (5), with the main difference that in that study the nutrient input is prescribed with fixed values, while here it is determined by the flow via S b (t, x, y). Another important point is that in this work the ecological fields are initialized to constant values N n, P n and Z n all over the domain (Table 2). These values correspond to the equilibrium solution of the NPZ system, that is, those obtained as the system relaxes towards an equilibrium state when the only input of nutrients is given by S n (i.e. setting S b = ). Thus, in absence of additional sources or sinks of nutrients, the system evolves towards these constant, normal conditions.

L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 261 Escarpment Seamount 2 2 B A 3 y profile 3 y profile 4 4 Depth (m) 4 Depth (m) 4 Fig. 1. Surfaces and lateral profiles of the two bottom topographies adopted in this study [Eqs. (1) and (12)]: (Left) escarpment, (Right) submarine mountain. Gray (white) surfaces correspond with shallow (deep) regions. In both cases, the maximum (minimum) depth is m (4 m). Time series of the NPZ fields are calculated at points A, B, L and R, as discussed in Sections 3 and 4. Table 2 Parameters used in the NPZ model. Symbol Value β.66 mmol N m 3 γ.75 ɛ 1. (mmol N m 3 ) 2 d 1 a 2. d 1 k N.5 mmol N m 3 µ N.2 µ P.3 d 1 µ Z.2 (mmol N m 3 ) 1 S n N n P n Z n 3. Escarpment.65 d 1.238 mmol N m 3.3796 mmol N m 3.4 mmol N m 3 Before discussing the dynamics of the vortex and of the NPZ fields on the escarpment, it is necessary to briefly describe the basic vortex behavior in the presence of a topographic slope. It is also important to emphasize the difference between the flow evolution when the scale of the vortex is smaller than the horizontal scale of the topography, and the behavior when these two scales are of the same order. Consider first an f plane in which the topography is a weakly sloping bottom. A cyclonic vortex on this plane drifts uphill with a leftward inclination (looking upslope), due to relative vorticity redistribution at the right (left) flank of the vortex as fluid moves uphill (downhill) (see e.g. Carnevale et al., 1991). This behavior is dynamically equivalent to the northwest drift of cyclones on a β-plane, with the shallow (deep) regions dynamically equivalent to the geographical north (south). Such a configuration is usually called a topographic β-plane, and it is often used in laboratory experiments to simulate the drift of geophysical vortices (see a recent review by van Heijst and Clercx (9)). The topographic parameter β b, which plays a role that is equivalent to the latitudinal variation of the Coriolis parameter in large-scale geophysical flows, is given by β b = αf /H, where α is the bottom slope and H the mean flow depth. The motion of a vortex in this system implies the generation of topographic Rossby waves. These oscillations have a scale that is larger than that of the vortex, and are characterized by a series of opposite-sign circulation cells behind the cyclone path, that rapidly occupy the whole domain (of course, a similar pattern occurs with an anticyclone). On the other hand, a topographic escarpment refers to a much steeper sloping bottom. The evolution of a vortex over this type of topography is characterized by a fast decomposition of the flow into topographic waves due to the much stronger β b (Grimshaw et al., 1994; Zavala Sansón and van Heijst, ). The aim in this section is to examine the evolution of the ecological NPZ fields affected by the presence of topographic waves generated over a steep escarpment. It must be emphasized that the vertical motion associated with up and downhill displacements of fluid columns represents a source or sink of nutrients in the NPZ model, and therefore plankton populations will be eventually affected.

262 L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 2 Relative vorticity 2 Divergence t = d t = d 2 2 Fig. 2. Contours of a cyclonic vortex in presence of an escarpment at t =. Left: Relative vorticity (contour interval is ω /1 = 3.68 1 6 s 1 ). Right: Corresponding contours of the divergence field with an interval of.77 d 1. Positive (negative) contours are plotted with black (gray) lines. The topography is represented by the gray-scale surface, where white indicates the deeper part. In order to represent the escarpment, a slope oriented in the y direction is adopted h(y) = H s h s tanh[(y y c )/W], (1) where W = 2 km is the horizontal scale of the escarpment, the mean depth is H s = 4 m, h s = m, y c = 128 km. Note that h = m at the southern side of the domain, whilst h = 4 m at the other side of the escarpment. The corresponding topographic parameter is β b = h s f /H s W 5 1 1 (ms) 1 ; the slope in this case is α = h s /W. The dispersion relation for topographic waves along the x direction is σ = β bk (11) W 2 + k 2 where σ is the frequency and k the wavenumber. This relation indicates that a perturbation with wavelength km has a period of about 1 days, i.e. a phase speed of about 1 km/day. The flow is set in motion by the cyclonic vortex located at the deeper side of the slope. Fig. 2 shows the initial vorticity distribution and the corresponding divergence field produced by the vortex. Note that the divergence has non-zero values only over the slope because fluid columns are stretched and squeezed only over this region. There are two regions of opposite sign: at the right side of the vortex divergence is positive because there is upwelling as fluid columns move towards shallower waters. In contrast, divergence is negative at the left side of the cyclone, where fluid columns are stretched as they move downhill. Of course, this will imply an input (output) of nutrients at the right (left) flank of the vortex. From its initial position, the vortex drifts towards the topography and to the left, according to the topographic β effect, and it is rapidly decomposed into alternate cells of positive and negative vorticity traveling westward, that is, with shallow water to the right. The important point to illustrate here is the evolution of the NPZ fields over the slope, associated with the generation of topographic waves. Fig. 3 shows the divergence and the NPZ fields at day 8, when the trail of topographic Rossby cells is well developed. These structures can be observed from the divergence field, where the regions of upwelling and downwelling are clearly formed (first panel). A similar wave pattern, although with a more complicated structure, is observed for phytoplankton and zooplankton abundance, as shown in the lower panels. A first conclusion is, therefore, that the characteristic behavior of the ecological model over an escarpment is the development of regions with alternate rich and poor amounts of nutrients and plankton biomass. The biological fields and the horizontal divergence do not have exactly the same structure and, more importantly, the same phase (their maxima do not coincide). A useful way to observe this is by considering the time series of NPZ and divergence fields at specific points over the escarpment. This is done for the two points marked with a cross in Fig. 1: one is at the right side of the domain (A) while the other is at the center (B). Both locations are situated exactly in the middle of the escarpment (in the y direction) and their depth is 4 m. Fig. 4 shows the corresponding 3-days time series of the NPZ fields (upper panels) and the calculated divergence (lower panels). Consider first the time series of horizontal divergence (S b ) and nutrients at point A. The time evolution of local divergence shows a very clear co-sinusoidal pattern with period of about 11 days. This value corresponds quite well to a topographic perturbation with a wavelength of about km or slightly less, as obtained by the dispersion relation. The maximum values of S b are of the order of.35 d 1, which is about 2 times larger than the normal relaxation rate (S n =.65 d 1 ). The amount of nutrients shows a rapid response, increasing when there is positive divergence and decreasing when the flow converges. Note that nutrients in the surface layer can be completely depleted during periods of strong negative divergence (say, between days 5 8 and days 17 19, approximately). The oscillatory character is also observed in the phytoplankton and zooplankton series. The available nutrients, and therefore the concentrations of plankton populations, are determined by the oscillatory behavior of the horizontal divergence at this point. Advection effects turn out to be of limited relevance or, in other words, the NPZ fields are not swept away by the flow. This can be proven by solving the NPZ system in homogeneous conditions, using a time-dependent prescribed rate S b (t) = S cos(2πt/t). Values for the amplitude and period of this periodic input are S =.35 d 1 and T = 11 d, respectively. Even with this crude approximation of the topographic wave signal, the resulting time series (not shown here) are very similar to those in Fig. 4: the values obtained in this way differ by at most 1% from those provided by the coupled model. The main factor in the dynamics of the NPZ fields is therefore the oscillatory character of the topographic waves. Direct advection of ecological fields is of limited importance here, as the waves do not transport water masses. In next section we show that advection effects play a more important role in the dynamics of biological fields over a topographic seamount. The shape of the signal at point A is quite persistent when different biological parameters are used (see Table 2). Simulations using larger values of β and ɛ, for instance, have shown similar

L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 263 Divergence N 2 2 P Z 2 2 Fig. 3. Contours of horizontal divergence and NPZ fields over an escarpment at t = 8 days (two vortex rotation periods): (a) Divergence (contour interval as in Fig. 2); (b) N N n (N n /1 =.238 mmol N m 3 ); (c) P P n (P n /1 =.3796 mmol N m 3 ); (d) Z Z n (Z n /1 =.4 mmol N m 3 ). Positive (negative) contours are plotted with black (gray) lines. For the biological fields we have plotted the anomalies from the equilibrium values N n, P n, Z n. 1 Point A 1 Point B N,P,Z (mmol Nm 3 ). 8. 6. 4. 2 N,P,Z (mmol Nm 3 ).8.6.4.2 1 2 3 1 2 3.4.1 Divergence (d 1 ).2.2 Divergence (d 1 ).5.5.4 1 2 3.1 1 2 3 Fig. 4. Time series of NPZ fields (upper panels) and horizontal divergence (lower panels) at points A = ( km, 128 km) and B = (128 km, 128 km), both located over the escarpment (at a local depth of 4 m). N: solid line, P: dashed line, Z: dashed dotted line.

264 L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 Relative vorticity t =4 d Divergence t =12 d t =4 d t =12 d Fig. 5. Contours of a cyclonic vortex over a submarine mountain at t = 4, 8, 12 days. First row: relative vorticity. Second row: divergence field. Contour intervals as in Fig. 2. trends of the time series. Both cases correspond to higher consumption rates for phytoplankton and zooplankton, respectively. The main point is that plankton abundance is affected only in amplitude but not in relative phase. Time series at point B have a less coherent shape, as shown in the corresponding panels of Fig. 4. This is due to the nonlinear evolution and dispersive character of the waves over the slope. However, the response of nutrients to the horizontal divergence remains similar: positive (negative) divergence increases (decreases) the amount of nutrients. Phytoplankton and zooplankton abundances also show a similar lag as at point A. 4. Submarine mountain In this section we explore the behavior of the NPZ system under the influence of a cyclonic vortex moving over a submarine mountain. The fluid depth associated with the seamount has the following form: h(x, y) = H m h m exp(r 2 /M 2 ), (12) where r = x 2 + y 2 is the radial distance from the center of the domain, H m = m is the maximum depth, M = 2 km and h m = m are the horizontal scale and the mountain height, respectively. The shallowest point on top of the mountain is 4 m. The physical dimensions of the topography selected here resemble those in a relatively shallow system as the El Bajo de Espiritu Santo seamount in the southern Gulf of California (Trasviña-Castro et al., 3) or the upper active layer of the ocean at the Great Meteor seamount (Beckman and Mohn, 2) and the Minamikasuga seamount studied by Genin and Boehlert (1985). The initial condition is the same cyclone as for the case of the escarpment, starting at the same position. The evolution of cyclonic vortices over a conical bottom has been studied by Carnevale et al. (1991). Initially, the vortex moves uphill and to the left, according to the associated topographic β-effect due to the nearest slope; as the vortex impinges over the topography it turns anticyclonically around the top of the mountain. This behavior is explained by the fact that the local orientation of the slope (responsible for the topographic β) is continuously changing as the vortex drifts. Fig. 5 shows the calculated relative vorticity contours during 3 turnover periods of the vortex, after which it makes about half a revolution around the mountain. During the whole simulation (3 days) the cyclone makes approximately 1.5 clockwise revolutions. However, the shape of the vortex is not circular, but elongated. In fact, the deformed contours show two relative maxima at day 8. Another important feature is the creation of anticyclonic vorticity over the top of the mountain that remains there during the whole simulation. This well-known behavior is expected as the flow tends to acquire negative vorticity over shallow regions (where fluid columns are squeezed). The divergence field, shown in Fig. 5 for the same times (lower panels), is characterized by a dipolar structure as in the case of a bottom slope, but in this case it is always centered on top of the mountain, even when the vortex is initially separated from it. It is important to notice that these up- and downwelling areas rotate as a whole in clockwise sense, according to the motion of the vortex. Note the coherence of the divergence field over the top of the mountain, that contrasts with the strong distortion of the vortex. The dipolar structure of the divergence is a manifestation of the so-called trapped waves (e.g. Beckman and Mohn, 2). The NPZ fields are shown in Fig. 6 for the same times. Nutrient concentration is strongly increased over the mountain, while maintaining a low-value zone around it. This nutrientrich area remains over the top of the seamount during most of the simulation, despite the fact that the divergence field shows positive and negative values around this region. In other words, as long as the vortex remains around the topographic feature,

L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 265 N t =4 d t =12 d P t =4 d t =12 d Z t =4 d t =12 d Fig. 6. Contours of NPZ fields over a submarine mountain at t = 4, 8, 12 days (contour intervals as in Fig. 3). First row: N N n ; Second row: P P n ; Third row: Z Z n. Contour intervals as in Fig. 3. nutrients are abundant on top of it. This in turn implies a favorable environment for phytoplankton and zooplankton, whose populations are also increased over the seamount. Time series of divergence and NPZ fields at two specific points are shown in Fig. 7. The two points L and R are located 18 km to the left and right from the summit of the mountain, respectively (see Fig. 1). The divergence field induces a very clear signal: at point R, divergence is positive during about 1 days, then it becomes negative and at the end it turns positive again. This behavior corresponds to the clockwise motion of the dipolar structure of the divergence field. A similar trend is found at point L, but shifted in time of about 1 days. At a first glance, it seems that nutrients in the surface layer are increased or decreased according to horizontal divergence, as in the case of the escarpment. The behavior, however, as well as the temporal evolution of plankton concentrations, is more complicated. This can be illustrated when considering the following two observations for point L: (a) during days 4 to 1 there is an increase of nutrients, while divergence is negative, and (b) phytoplankton and zooplankton are suddenly increased simultaneously, from day 4 to 5, approximately. The reason for these two behaviors is the advection of biological fields as the flow turns around the mountain, influencing the local values. Similar arguments can be applied to almost the whole series at points L and R. To demonstrate that advection is responsible for this behavior, we have solved the NPZ system in homogeneous conditions with an appropriate time-varying nutrient input (obtained by imposing a time-dependent rate S b (t) of the same form of the divergence curve in lower panels of Fig. 7), as done for the case of the escarpment. We have verified that the result is completely different from that shown in upper panels of Fig. 7. This indicates that the dynamical evolution of the flow over the mountain is a fundamental factor or, in other words, that the NPZ system must be coupled with a dynamical model. 5. Discussion and conclusions The dynamics of nutrients, phytoplankton and zooplankton at the ocean surface has been analyzed by coupling an NPZ formulation with a physical quasi two-dimensional flow model

266 L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 1 Point L 1 Point R N,P,Z (mmol Nm 3 ).8.6.4.2 N,P,Z (mmol Nm 3 ).8.6.4.2 1 2 3 1 2 3 Divergence (d 1 ).6.4.2.2.4 1 2 3 Divergence (d 1 ). 1.5.5.1 1 2 3 Fig. 7. Time series of NPZ fields (upper panels) and horizontal divergence (lower panels) at points L = (11 km, 128 km) and R = (146 km, 128 km), located at the leftand right-hand sides of the seamount, respectively. N: solid line, P: dashed line, Z: dashed dotted line. with topography. The main goal has been to examine the evolution of the ecological fields as a mesoscale flow (namely a cyclonic vortex) encounters two specific topographic features: a steep escarpment and a submarine mountain. Coupling of physical and biological models is made by using the horizontal divergence of the flow calculated in the dynamical model, resulting from the interaction with the topography, as a source/sink term in the equation for nutrients in the NPZ formulation. The results show the potential of mesoscale phenomena related with bottom topography to affect the behavior of plankton ecosystems. The main dynamical response of the flow over the escarpment is the generation of topographic Rossby waves, which travel along the topographic contours with shallow water to the right. These waves are characterized by alternate cells of relative vorticity with opposite sign. The resulting horizontal divergence field S b has a similar form, as the flow moves up and downhill, stretching and squeezing fluid columns. The response of the nutrient field N shows a similar pattern, since horizontal divergence adds or subtracts nutrients according to its sign and magnitude (see Fig. 3). This was also clearly shown in Fig. 4, where time series of S b and N are measured at single points over the escarpment. Phytoplankton and zooplankton fields have a different spatial structure due to their different temporal response to available resources; phytoplankton have a growth timescale of about 3 days, while for zooplankton the characteristic growth time is about 8 days, which is reflected in the phase shift observed in the time series of these fields. Another important result is that the main structure of the biological time series remains the same (except in amplitude) when varying the main parameters of the NPZ system. These observations suggest that point measurements of ecological fields over steep slopes, such as shelf breaks, could reflect the characteristics of topographic submesoscale and mesoscale oscillations. In other words, the time series of such measurements should take into account the expected periods of topographic waves. Recall that, in the present case, an escarpment that is 2 km wide implies a topographic wave with a period of about 11 days and horizontal length scale of about km. Of course, in realistic situations such a signal can be expected to be highly reduced or contaminated by the presence of other relevant effects ignored here: stratification effects, additional currents and/or biological mechanisms not explicitly considered in the present model. Nevertheless, the results presented here are quite robust and remain as an open subject to be further investigated with in-situ data. The results observed for the topographic seamount are very different. Close to a seamount, cyclonic vortices perform clockwise revolutions around the top, where an anticyclonic structure is generated. The divergence field displays a clockwise-rotating dipolar structure centered at the summit of the mountain, while the nutrient field shows a remarkable increase at this position (about 5 to 7 times the original value, N n ). Phytoplankton and zooplankton populations are also enhanced but their maxima are off the center of the seamount, and the magnitude of this increment is much less dramatic (about twice their original reference values, P n and Z n ). Note that nutrients are accumulated over the top of the seamount already in the early stages of the simulation, when the vortex approaches and impinges on the mountain. As the flow evolves, the amount of these injected nutrients remains approximately stable. Some authors (e.g. Beckman and Mohn, 2; Genin, 4) have pointed out two important conditions that upwelled nutrients due to a topographic feature must satisfy in order to increase phytoplankton and zooplankton at the sea surface: (1) upwelling must be strong enough to reach the surface layer, and (2) it must have a sufficiently long duration, in order to allow phytoplankton and zooplankton to grow. Some field studies on seamounts in the Pacific ocean indicated that there is only sparse and not conclusive evidence supporting the relevance of upwelling on biological aggregations over these topographic features (Genin, 4). Indeed, upwelled nutrient-rich waters might be advected away from the mountain on very short timescales (of a few hours or days), not allowing for the development of phytoplankton, and hence of zooplankton. The present study, however, albeit limited to idealized conditions, shows that the upwelling mechanism is sufficiently

L. Zavala Sansón, A. Provenzale / Theoretical Population Biology 76 (9) 258 267 267 strong and persistent, so that the NPZ fields are significantly enhanced over the top of a submarine mountain during periods of several days. An important difference between the two topographies examined here is related to the role of advection. The dynamics of biological fields over an escarpment is only weakly affected by advection, while horizontal transport effects play a major role over a seamount. The dynamics of the NPZ fields at one point on the escarpment is governed by the passage of topographic waves, which do not transport water masses. In the case of the mountain, by contrast, perturbations triggered by the vortex interfere with each other as they propagate around the summit. In this case, the vortex, a strongly nonlinear feature with closed streamlines, can transport water and biological fields over significant distances. This provides an indication on the complexity of the NPZ fields over seamounts, even in the simplified configuration used here. It also emphasizes a similar conclusion by Beckmann and Mohn (2), who pointed out that interpretations of biological data based on Eulerian time series might be inadequate. Thus, one can expect even greater uncertainties when interpreting biological data for non circular topographies or more irregular seamounts, as underlined by Trasviña-Castro et al. (3). Before closing, we stress some of the assumptions and drawbacks implicit in our formulation of phytoplankton dynamics. In our approach, the biological fields are governed by one of the simplest versions of the NPZ model, which has sometimes been criticized for its over-simplifying assumptions (see the comments by Franks (2)). Besides, the phytoplankton s response to light is ignored. These choices have been made for keeping the system as simple as possible, in order to elucidate the basic effects of the interaction between topography-induced vertical velocity and marine ecosystem dynamics. Future works will have to consider a more realistic and complex marine ecosystem model. Note, also, that we have chosen to work only with reaction-advection terms in the NPZ model, neglecting the diffusion of biological fields. Given the spatial scale of the fluid parcels (of the order of kilometers), and the temporal size of the time steps (hours), the effects of smallscale turbulent diffusion are largely insignificant, as discussed in detail in Pasquero et al. (5). On these spatial scales the diffusive time scale is much longer than the time scales associated with both the advection and the reactions affecting the plankton concentrations. On the other hand, a significant advantage of the formulation used here is the method that we use to couple the biological and physical models, based on using the value of the horizontal divergence that is directly computed from the flowtopography interaction. This procedure is different from what is usually done when using a two-dimensional or a quasigeostrophic model. In contrast with those formulations, the shallow water approximation adopted here allows for an explicit expression of the horizontal divergence and hence of the vertical velocity field, at any time and spatial position of the flow. This fact allows for a more natural description of the input/output of nutrients in the upper layer of the ocean, avoiding artificial parameterizations of the spatial and temporal variations of the upwelling and downwelling processes. References Beckman, A., Mohn, C., 2. The upper ocean circulation at the Great Meteor seamount. Part II: Retention potential of the seamount-induced circulation. Oc. Dyn. 52, 194 24. Carnevale, G.F., Kloosterziel, R.C., van Heijst, G.J.F., 1991. Propagation of barotropic vortices over topography in a rotating tank. J. Fluid Mech. 233, 119 139. Comeau, L.A., Vézina, A.F., Bourgeois, M., Juniper, S.K., 1995. Relationship between phytoplankton production and the physical structure of the water column near Cobb seamount, northeast Pacific. Deep Sea Res. 42, 993 5. Franks, P.J.S., 2. NPZ models of plankton dynamics: Their construction, coupling to physics, and application. J. Oceanogr. 58, 379 387. Genin, A., 4. Bio-physical coupling in the formation of zooplankton and fish aggregations over abrupt topographies. J. Mar. Sys., 3 2. Genin, A., Boehlert, G.W., 1985. Dynamics of temperature and chlorophyll structures above a seamount: An oceanic experiment. J. Mar. Res. 43, 97 924. Grimshaw, R., Tang, Y., Broutman, D., 1994. The effect of vortex stretching on the evolution of barotropic eddies over a topographic slope. Geophys. Astrophys. Fluid Dyn. 76, 43 71. Huppert, H.E., Bryan, K., 1976. Topographically generated eddies. Deep Sea Res. 23, 655 679. Koszalka, I., Bracco, A., Pasquero, C., Provenzale, A., 7. Plankton cycles disguised by turbulent advection. Theo. Pop. Biol. 72, 1 6. Lévy, M., Klein, P., Tréguier, A.M., 1. Impacts of submesoscale physics on phytoplankton production and subduction. J. Mar. Res. 59, 535 565. Lévy, M., 3. Mesoscale variability of phytoplankton and of new production: Impact of the largescale nutrient distribution. J. Geophys. Res. 18, 3358. Lévy, M., 8. The modulation of biological production by oceanic mesoscale turbulence. In: Weiss, J.B., Provenzale, A. (Eds.), Transport and Mixing in Geophysical Flows. Springer, Berlin, pp. 219 261. Mann, K.H., Lazier, J.R.N., 6. Dynamics of Marine Ecosystems: Biological-Physical Interactions in the Ocean. Blackwell Pub, 496 pp. Martin, A.P., Richards, K.J., Bracco, A., Provenzale, A., 2. Patchy productivity in the open ocean. Global Biogeochem. Cycles 16, 125. McGillicuddy, D.J., Robinson, A.R., 1997. Eddy-induced nutrient supply and new production in the Sargasso Sea. Deep Sea Res. Part I 44, 1427 14. Moulins, A., Rosso, M., Ballardini, M., Wurtz, M., 8. Partitioning of the Pelagos Sanctuary (north-western Mediterranean Sea) into hotspots and coldspots of cetaceans distributions. J. Mar. Biol. Assoc. U.K. 88, 1273 1281. Oey, L.C., Lee, H.C., 2. Deep eddy energy and topographic Rossby waves in the Gulf of Mexico. J. Phys Oceanogr. 32, 3499 3527. Pasquero, C., Bracco, A., Provenzale, A., 5. Impact of spatiotemporal variability of the nutrient flux on primary productivity in the ocean. J. Geophys. Res. 11, C75. Trasviña-Castro, A., Gutierrez de Velasco, G., Valle-Levinson, A., González-Armas, R., Muhlia, A., Cosio, M.A., 3. Hydrographic observations of the flow in the vicinity of a shallow seamount top in the Gulf of California. Est. Coast. Shelf Sci. 57, 149 162. Uz, B.M., Yoder, J.A., Osychny, V., 1. Pumping of nutrients to ocean surface waters by the action of propagating planetary waves. Nature 49, 597 6. van Heijst, G.J.F., Clercx, H.J.H., 9. Laboratory modeling of geophysical vortices. Annu. Rev. Fluid Mech. 41, 143 164. Zavala Sansón, L., van Heijst, G.J.F.,. Interaction of barotropic vortices with coastal topographies: Laboratory experiments and numerical simulations. J. Phys. Oceanogr. 3, 2141 2162. Zavala Sansón, L., van Heijst, G.J.F., 2. Ekman effects in a rotating flow over bottom topography. J. Fluid Mech. 471, 239 256.