You are on page 1of 73

SPRINGER BRIEFS IN ENVIRONMENTAL SCIENCE

Eduardo Marcelo Acha
Alberto Piola
Oscar Iribarne
Hermes Mianzan

Ecological
Processes at
Marine Fronts
Oases in the Ocean
SpringerBriefs in Environmental Science
SpringerBriefs in Environmental Science present concise summaries of cutting-
edge research and practical applications across a wide spectrum of environmental
fields, with fast turnaround time to publication. Featuring compact volumes of 50
to 125 pages, the series covers a range of content from professional to academic.
Monographs of new material are considered for the SpringerBriefs in Environmen-
tal Science series.
Typical topics might include: a timely report of state-of-the-art analytical tech­
niques, a bridge between new research results, as published in journal articles and
a contextual literature review, a snapshot of a hot or emerging topic, an in-depth
case study or technical example, a presentation of core concepts that students must
understand in order to make independent contributions, best practices or protocols
to be followed, a series of short case studies/debates highlighting a specific angle.
SpringerBriefs in Environmental Science allow authors to present their ideas and
readers to absorb them with minimal time investment. Both solicited and unsolicited
manuscripts are considered for publication.

More information about this series at http://www.springer.com/series/8868


Eduardo Marcelo Acha · Alberto Piola
Oscar Iribarne · Hermes Mianzan

Ecological Processes
at Marine Fronts
Oases in the Ocean

13
Eduardo Marcelo Acha and
Instituto Nacional de Investigación
y Desarrollo Pesquero (INIDEP) Consejo Nacional de Investigaciones
Proyecto Ecología Pesquera Científicas y Técnicas
Mar del Plata Buenos Aires
Argentina Argentina

and
Oscar Iribarne
Instituto de Investigaciones Marinas Instituto de Investigaciones Marinas
y Costeras (IIMyC) y Costeras (IIMyC)
Consejo Nacional de Investigaciones Científicas Consejo Nacional de Investigaciones Científicas
y Técnicas-Universidad Nacional de Mar del Plata y Técnicas-Universidad Nacional de Mar del Plata
Mar del Plata Mar del Plata
Argentina Argentina

Alberto Piola Hermes Mianzan


Departamento Oceanografía Instituto Nacional de Investigación
Servício de Hídrografia Naval y Desarrollo Pesquero (INIDEP)
Buenos Aires Proyecto Ecología Pesquera
Argentina Mar del Plata
Argentina
and
and
Departamento de Ciencias de la Atmósfera
y los Océanos, Facultad de Ciencias Instituto de Investigaciones Marinas
Exactas y Naturales y Costeras (IIMyC)
Universidad de Buenos Aires Consejo Nacional de Investigaciones Científicas
Buenos Aires y Técnicas-Universidad Nacional de Mar del Plata
Argentina Mar del Plata
Argentina
and

Instituto Franco-Argentino Sobre Estudios


del Clima y sus Impactos
Buenos Aires
Argentina

Hermes Mianzan is deceased

ISSN  2191-5547 ISSN  2191-5555  (electronic)


SpringerBriefs in Environmental Science
ISBN 978-3-319-15478-7 ISBN 978-3-319-15479-4  (eBook)
DOI 10.1007/978-3-319-15479-4

Library of Congress Control Number: 2015930817

Springer Cham Heidelberg New York Dordrecht London


© The Author(s) 2015
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction
on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and
regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed
to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty,
express or implied, with respect to the material contained herein or for any errors or omissions that may have been made.

Printed on acid-free paper

Springer International Publishing AG Switzerland is part of Springer Science+Business Media


(www.springer.com)
We dedicate this work to the memory of our
colleague, friend, and co-author, Hermes
W. Mianzan, who passed away during the
final editorial steps. Hermes was an
insightful researcher, and an enthusiastic
teacher of the ecology of fronts. Endowed
with a perennial sense of humor, he once
told his students that “… the only important
things are fronts. The rest is just water for
fishes to travel from one front to another.”

Eduardo Marcelo Acha


Alberto Piola
Oscar Iribarne
Contents

1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Frontal Types. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1 Tidal Fronts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Shelf-Break Fronts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Upwelling Fronts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Estuarine Fronts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.5 Plume Fronts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6 Fronts Associated with the Convergence or Divergence
of Water Masses at High Seas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.7 Frontal Eddies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.8 Fronts Associated to Geomorphic Features. . . . . . . . . . . . . . . . . . . . 12

3 Biology of Fronts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1 Biological Production. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 Trophic Webs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3 Biogeography. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4 Diversity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Life Histories Traits in Relation to Fronts. . . . . . . . . . . . . . . . . . . . . 26
3.6 Migrations and Transport. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4 Management and Conservation of Marine Life. . . . . . . . . . . . . . . . . . . 33


4.1 Fisheries. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2 Conservation Issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.3 Climate Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5 Comparisons of Fronts with Other Boundaries at Sea. . . . . . . . . . . . . 41


5.1 The Pycnocline Interface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2 The Sea Water-Sediment Interface. . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 The Sea Surface-Atmosphere Interface. . . . . . . . . . . . . . . . . . . . . . . 44

vii
viii Contents

5.4 The Ice-Water Interface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45


5.5 Fronts Contrasted with the Other Interfaces. . . . . . . . . . . . . . . . . . . . 46

6 Comparisons of Fronts with Terrestrial Boundaries


and the “Ecotone” Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

7 Final Remarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.1 Landmarks and Beacons: Orienting and Meeting . . . . . . . . . . . . . . . 54
7.2 Mechanical Energy for Retention. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.3 Mechanical Energy for Biological Production. . . . . . . . . . . . . . . . . . 56

Literature Cited. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Chapter 1
Introduction

On almost any ocean there is day after day when scarcely


a bird, beast, or fish appears to break the monotony, and
contrasted with this, here and there a few minutes or hours of
teeming abundance, when every field of the binoculars shows a
dozen birds or a hundred, and porpoises, orcas, or that cyno-
sure of all passengers’ eyes, “real whales,” are seen on every
hand. Why should things be so badly arranged?
(Brooks 1934)

Abstract Marine fronts are part of the structural complexity of the sea; they
are narrow boundaries separating different water masses. Fronts are caused by
diverse forcing and occur throughout the world ocean at several spatial and tem-
poral scales. Unlike terrestrial ecotones, they show high biological production,
affecting pelagic and benthic organisms of all trophic levels; consequently they
are important for fisheries. Solar energy stimulates biological production in the
entire biosphere, but in the sea it needs to be complemented by auxiliary energy
to replenish plant nutrients; a significant quantity of mechanical energy becomes
available for biological production at fronts. Global change is redistributing auxil-
iary energy in the oceans; consequently fronts are ideal sites for early monitoring
of global change effects. Fronts also provide retention mechanisms for plankton
in the highly dispersive marine environment; and become landmarks and bea-
cons important for migrations or meeting of some species in a traceless realm.

Keywords Biological production · Plankton retention · Climate change · 


Migrations  ·  Ecological boundaries  · Ecotones

Though not immediately perceived, oceans are characterized by a profusion of


patterns which cover several scales of space and time. A few decades ago Le Févre
(1986) began his review on shelf tidal fronts stressing that “A common-sense view
of the marine environment as a fluid medium would probably imply progressive
changes and smooth gradients in physical properties. Sharp boundaries, however,

© The Author(s) 2015 1


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_1
2 1 Introduction

are actually quite widespread and are generally known as fronts…”. Increased
availability of satellite imagery nowadays challenge such a common-sense view
of the oceans as being a nearly homogeneous habitat, displaying at a glance that
the marine environment is characterized by very complex structures (Fig. 1.1). The
large-scale views of the ocean’s surface characteristics from earth-orbiting satel-
lites, together with high resolution field measurements and numerical simulations,
have transformed our perceptions of the range of scales and the variety of patterns
of the marine ecosystems (Bakun 1996). Regions in the ocean are organized by
temperature and salinity. Those variables do not change gradually with horizon-
tal distance; instead there are large regions where horizontal gradients are small,
bounded by narrow regions (referred to as fronts) where horizontal gradients are
high. Fronts localize at the meeting of two different water masses. At these bound-
aries, particular biological and ecological processes occur that are more deter-
minant of the ecological properties of the region than the phenomena occurring
inside both water masses (Frontier 1986). Fronts are an integral part of the sea, of
its fluid processes and of its ecological functioning, and are characterized by high
biological activity (Cushman Murphy 1944; Sournia 1994; Acha et al. 2004).
But fronts are not merely a meeting of waters with dissimilar properties, more
importantly, they are a dynamic phenomenon. They are vertically inclined inter-
faces between different water masses, and frequently present a rather complex
three-dimensional structure (Uda 1938; Fedorov 1986). Fronts occur on a variety of
length scales, from a few meters up to many thousands of kilometers. They can be
short-lived (days), although many fronts are quasi-stationary and seasonally persis-
tent; prominent fronts are present year-around. Cross-frontal differences in SST (sea
surface temperature) and sea surface salinity can be as large as 10–15 °C and 2–3

Fig. 1.1  Sea surface
temperature satellite image
of the Brazil-Malvinas
Confluence region. Note
the highly structured spatial
pattern. Image processed
at the Rosenstiel School of
Marine and Atmopsheric
Science, University of
Miami, USA
1 Introduction 3

practical salinity units (psu, we use dimensionless salinity hereafter), respectively;


typical differences are 2–5 °C and 0.3–1.0 psu. The vertical extents of fronts vary
from a few meters to more than a kilometer, with some major fronts reaching the
open ocean bottom at depths exceeding 4 km (Belkin et al. 2009). Circulation at
the fronts is usually associated with a density difference between the two meeting
waters, which can drive relatively intense (geostrophic) flows in the along-front
direction. Fronts are also sustained by convergence at the surface or bottom bound-
ary, approximating an interface, even in the presence of diffusive effects (Largier
1993). Unlike other oceanographic phenomena, marine fronts are physically tan-
gible. Fronts are characterized by surface manifestations such as lines of rips and
foam, rough water zones, accumulations of floating objects, marked changes in
water color, and high biological activity (Fig. 1.2). They are not abstract demarca-
tion boundaries between equally abstract water masses, or mere tightening of iso-
lines, but real, complex and varied dynamic phenomena (Uda 1938; Fedorov 1986).
Fronts have been noted by fishermen, early voyagers and marine scientists cen-
turies ago (Box 1). Though the name “front” was not in use, they were recorded
in the literature for over two centuries taking note of their most evident character-
istics: the profusion of life and the contrasting features (color; roughness) of the

Fig. 1.2  Some fronts are characterized by surface manifestations such as marked changes in


water color and/or roughness. Fronts at a Mobile Bay, USA (L. Chiaverano), b Patos Lagoon
mouth, Brazil (O. Möller), c Río de la Plata (Argentina-Uruguay) (M. Framiñan), d meeting of
the Baltic and North seas (Denmark) (E. Spencer http://top10rate.com/top-10-spectacular-ocean-
phenomena/facts)
4 1 Introduction

meeting waters (Franklin 1786; Darwin 1845; Beebe 1926; Cushman Murphy
1944); it was also evident that some fronts were related to special bottom or coastal
configurations (Uda 1938). However, the ecological properties and functioning of
marine fronts has not been studied in detail until recently (Russell et al. 1999).

Box 1
Ancient reports of fronts: Outstanding biological activity attracted the
attention of ancient voyagers and scientists. In the past, this fact could have
been more notorious than nowadays because of a lesser impacted ocean
(fishing, contamination) and because of the use of slower and less disturbing
vessels (i.e. sailing ships).
In one day we passed through two spaces of water thus stained [by Cyanobacteria of
genus Trichodesmium], one of which alone must have extended over several square
miles… The line where the red and blue water joined was distinctly defined. The weather
for some days had been calm, and the ocean abounded, to an unusual degree, with liv-
ing creatures… how do the various bodies which form the bands with defined edges keep
together? …what causes the length and narrowness of the bands? … We must believe that
the various organized bodies are produced in certain favourable places, and are thence
removed by the set of either wind or water
(Darwin 1845)
The front moved past us, and we lay for some time in a turbulent area which contrasted
strongly with the preceding glassy calm … The surface was seething, boiling, with life,
much of which was de profundis. Larvae of clawless lobsters, tinted jellyfish, nurse chains
of salps, small herringlike fishes, a silvery hatchetfish with its face bitten off, rudder-fishes
hanging head downward, luminous lantern-fishes with shining light-pores, red and purple
swimming crabs, other creatures which we could not name at sight, and much that was too
small even to see distinctly—all swarming about under the searchlight, while pink squids,
from a few inches to a foot or more in length, kept darting from below, causing the show-
ers of fry from water to air and back again. A general holocaust was in progress. The little
fishes were eating invertebrates or straining out the plankton; the squids were pursuing
and capturing fishes of various sizes; and the blackfish were no doubt enjoying the squids
(Cushman Murphy 1944)
Chapter 2
Frontal Types

Abstract  Fronts are a dynamic phenomenon separating water masses of different


properties. They are narrow three-dimensional structures caused by diverse forcing
mechanisms; and are characterized by distinct physical, chemical, and biological
properties. Fronts occur throughout the world ocean at several spatial and tempo-
ral scales. Most stable fronts are steered by bottom topography. The most studied
frontal types are tidal fronts, shelf-break fronts, upwelling fronts, estuarine fronts,
plume fronts, fronts generated by convergence or divergence of water masses,
frontal eddies and fronts associated with abrupt topographic features.

Keywords  Classification of fronts  ·  Physical forcing  ·  Haline fronts  · Thermal


fronts  ·  Spatio-temporal scales of fronts

Fronts are caused by diverse forcing such as tides, continental run-off, conver-
gence of currents, wind, solar heating, bathymetry, etc. As a general rule, a front
in one property (e.g. temperature) can be detected in other properties. The con-
current physical, chemical, and biological manifestations of the same front are
typically collocated (Belkin et al. 2009). Long-term mean annual, seasonal, and
monthly frontal frequency maps for the Atlantic, Indian, and Pacific Oceans reveal
elevated concentrations of quasi-stationary fronts in coastal and marginal seas over
the entire world ocean between 75°N and 75°S (Fig. 2.1). Most of stable fronts are
steered by ocean bottom topography; the shelf break and upper continental slope
play the most important role in stabilizing their respective fronts (Belkin et al.
2009).
There is no a definitive classification of fronts, but a partial listing of them
would include tidal fronts, shelf-break fronts, upwelling fronts, estuarine fronts,
plume fronts, fronts associated with the convergence or divergence of water
masses in the open ocean, frontal eddies and fronts associated with geomorphic
features such as headlands, islands, and canyons (Mann and Lazier 2006).

© The Author(s) 2015 5


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_2
6 2  Frontal types

Fig. 2.1  Global distribution of maximum sea surface temperature gradient based on monthly


mean 0.25 × 0.25 degree resolution data from Reynolds et al. (2007). The map shows the larg-
est thermal fronts of the world ocean. BCU Benguela Current upwelling; CCU Canary Current
upwelling; CPU Chilean-Peruvian upwelling; CU California Current upwelling; EU Equatorial
upwelling; GSF Gulf Stream front; KF Kuroshio front; PF Polar front; SAF SubAntartic front;
SArF SubArtic front; STF SubTropical front

2.1 Tidal Fronts

In temperate climates, seasonal thermoclines develop near the surface during late
spring and summer. Unless some forcing provides enough mechanical energy to
mix the water column, thermoclines stabilize the water column, becoming stronger
as the warm season progresses. Tides are one of the main forcing processes in the
ocean (Munk and Wunsch 1998), whose energy distributes heterogeneously over
the global ocean. In continental shelves where a seasonal thermocline develops
and a high rate of tidal energy dissipation occur; there are regions in which the
intensity of turbulent mixing is able to continuously overcome the barrier to mix-
ing presented by stratification (Simpson and Hunter 1974; Pingree et al. 1975;
Le Févre 1986). As the tidal wave approaches the coast, the tidal amplitude and
its horizontal velocity gradually increase. At some critical depth, vertical turbu-
lence produced by friction between the tidal stream and the sea bed is sufficiently
enhanced (when added to turbulence produced by wind stress at the sea surface) as
to overcome the seasonal thermal stratification of the water column, giving rise to
tidally mixed regions near shore. Thus, the tidally mixed and the stratified regions
of the shelf are separated by a frontal region (Longhurst 1998). In contrast with
the more frequently observed situation described above there are some tidal fronts
where changes in mixing efficiency lead to a more strongly stratified water column
in the onshore region, such as in the southeast coast of Japan (Takeoka et al. 1997).
2.1  Tidal Fronts 7

Fig. 2.2  Frontal types. a Tidal front; b Shelfbreak front; c upwelling front (southern hemi­


sphere); d estuarine front. f Front; sf shelfbreak; t thermocline; w wind; arrows are currents;
dashed lines are lines of equal density

More research efforts have focused on tidal fronts than on any other frontal
types (Mann and Lazier 2006). Typical tidal fronts are seasonal and they estab-
lish each year at the same approximate time and location; and are characterized by
strong thermal gradients (Fig. 2.2a).

2.2 Shelf-Break Fronts

A linear zone of cool surface waters, supporting a plankton bloom, frequently


overlies the upper slope and shelf edge over continental shelves and other
banks elsewhere around the world ocean. Not all these fronts are forced by
the same mechanisms but the band of cold waters generally indicate entrain-
ment of deeper, cooler, and nutrient-rich waters towards the surface. A widely
accepted explanation of this observation involves the generation of internal
standing waves on the thermocline at the shelf edge where the tidal stream
encounters rough topography on the seabed (Longhurst 1998), but various
alternative explanations involving small-scale eddies (coupled with episodic
wind stress) have also been proposed. Additionally, the interleaving of water
masses at the front could enhance vertical stability, retaining phytoplankton
8 2  Frontal types

cells in the euphotic zone (Podestá 1990; Brandini et al. 2000). Numerical
simulations indicate that offshore flow in the bottom Ekman layer promotes
overturning over the continental shelf and detaches from the bottom at the
shelf break, where it mixes upward along sloping isopycnals, thus promoting
upwelling (Gawarkiewicz and Chapman 1992). High resolution hydrographic
observations across the Middle Atlantic Bight south of New England cor-
roborate the model results (Barth et al. 1998; Houghton and Visbeck 1998).
More recently Matano and Palma (2008) proposed a mechanism by which as a
downwelling current flows along the continental slope in the direction of con-
tinental trapped waves (e.g. with the coast on the left (right) in the Southern
(Northern) Hemisphere), bottom friction and lateral diffusion spread the flow
onto the neighboring shelf, thus generating along-shelf pressure gradients and
a cross-shelf divergence that is compensated by shelf-break upwelling. Though
shelf break fronts are topographically trapped and generally reorganize in a
few days after being disrupted (Gawarkiewicz and Chapman 1992), there are
strong indications of frontal instability that by enhancing cross-shelf exchange
might further promote nutrient enrichment. Despite of their ubiquitous occur-
rence, the fertilization mechanisms of shelf-break fronts seem to be diverse
(Fig. 2.2b).

2.3 Upwelling Fronts

Wind driven currents that flow towards the Equator along the western coasts
of continents (e.g. Peru; California; Benguela; Canaries Currents) are driven
away from the coasts due to the Earth’s rotation, leading to coastal upwelling
of nutrient rich waters. Upwelling fronts frequently present strong seasonality
derived from the seasonal variability of prevailing upwelling-favorable winds.
For instance along the coast of California the onset of upwelling occurs during
the “spring transition” when southerly winds reverse to upwelling-favorable, and
last until late fall (Huyer 1983). The sloping isopycnals sustain an along-shore
baroclinic jet. As the upwelling fronts are located a few tens of km from shore,
the interaction between the frontal jet and coastal indentations promote frontal
instabilities and vertical motions. These eastern boundary upwelling ecosystems
are among the most productive regions in the oceans (Pauly and Christensen
2005). The upwelled waters move away from the coast by Ekman transport and
converge at certain distance offshore, so the upwelled water sinks. Upwelling
fronts form at this interface between shelf water and the cool, nutrient-rich water
brought to the surface during wind-driven coastal upwelling. This frontal region
is highly productive and planktonic organisms aggregate on the coastal side of
the front and large numbers of fish concentrate at that location (Mann and Lazier
2006) (Fig. 2.2c).
2.4  Estuarine Fronts 9

2.4 Estuarine Fronts

Estuarine fronts are produced by the meeting of continental freshwaters and salty
marine waters. The later frequently form a salt-wedge below the former, leading to
the most frequently observed structure (Fig. 2.2d). These fronts develop usually in
bays, part of a bay, or inlets in which freshwater flows from land. These fronts are
controlled by salinity variations, and are frequently the most contrasting in terms
of water density. Estuarine fronts differ in stratification and dynamics mostly due
to diverse patterns in river discharge, and to the variable importance of external
forcing such as tides or wind. These frontal mechanisms lead to the formation of
plume, tidal intrusion and shear fronts at estuaries, some of which might develop
along-channel fronts, particularly during flood tide (O’Donnell 1993). Except for a
few very large estuaries (e.g. Río de la Plata; St. Lawrence) most estuarine fronts
have much smaller spatial scales than other types of marine fronts.
In some estuarine systems, a well-developed turbidity front characterizes the
innermost part of the estuary. This maximum gradient in turbidity is due to the
flocculation of suspended matter at the edge of the salt intrusion, and re-suspen-
sion of sediment due to tidal stirring. Turbidity fronts are clearly visible in satellite
images and frequently from the deck of ships (Fig. 2.2d).

2.5 Plume Fronts

In some situations, waters from either a river or an estuary pouring onto a con-
tinental shelf predominate over any tidal effects and flow into the neighboring
ocean creating a river plume, which may have a strong impact on the distribution
of water properties, sediments and biota. When the surface outflow onto the conti-
nental shelf is mainly of freshwater from the river itself, these plumes are referred
to as river plumes (e.g. those of the Mississippi or Amazon rivers); if the outflow
is of river waters mixed with salt water the flow constitutes an estuarine plume
(e.g. the Chesapeake; Río de la Plata or St. Lawrence estuaries). The Coriolis force
affects plumes turning them to the left (Southern Hemisphere) or right (Northern
Hemisphere), and the buoyant plume continues on its way as a coastal current par-
allel to the coast. Under favorable (downwelling) wind conditions buoyant river
plumes can extend hundreds of km away from the river mouth. Recent numerical
simulations have shown that in the absence of wind bottom-trapped plumes can
also propagate in the opposite direction (e.g. upstream). This is associated with
a baroclinic adjustment of the river discharge, while the downstream spreading
is generated by the cross-shelf barotropic pressure gradient (Matano and Palma
2010). Coastal currents; tidal currents and winds can modify plume dynamics in
a complex manner (Garvine 1975). At the boundary between the plume and the
marine coastal waters, the low salinity buoyant waters ride on the top of denser
saline waters, forming plume fronts where surface convergence and downwelling
occur (Mann and Lazier 2006) (Fig. 2.3a).
10 2  Frontal types

Fig. 2.3  Frontal types. a Plume front (southern hemisphere); b divergence front; c frontal eddies


(top panel) show eddies formation from a meandering current, w warm core eddy; c cold core
eddy. Lower panels show fronts (f) at convergences created in cyclonic or anticyclonic (south-
ern hemisphere) eddies, modified from Mann and Lazier (2006) and Bakun (2006a); d topo-
graphically controlled front (greenish areas indicate convergence frontal zones), modified from
Wolanski and Hamner (1988). f front; ss sea surface; t thermocline; w wind; arrows are currents;
dashed lines in a and b are lines of equal density

2.6 Fronts Associated with the Convergence or Divergence


of Water Masses at High Seas

In the Pacific, a westward current driven by the Trade winds is located roughly
between 5°S and 5°N. The Earth rotation leads to divergence of the Ekman layer
away from the equator, which is compensated by upwelling from subsurface layers
(Fig. 2.3b). The upwelling in turn is compensated by equatorward flows below the
mixed layers in both hemispheres (Wyrtki and Kilonsky 1984; Johnson et al. 2001).
Upwelling of cool subsurface water forms a cold tongue along the equator: the
equatorial upwelling and creates an extended thermal front of moderate intensity.
This thermal front shows some degree of seasonality in response to the seasonal
pattern of the trade winds (Mann and Lazier 2006). A relatively strong upwelling
system referred to as the Antarctic Divergence is observed in the Southern Ocean.
This system is caused by opposing southern hemisphere mid-latitude westerlies and
high-latitude easterlies. Here the winds and the Earth rotation drive a flow diver-
gence in the upper layers. Along the line of strongest wind stress curl separating
2.6  Fronts Associated with the Convergence … 11

the two wind systems the upper layer divergence is compensated by upwelling (the
Antarctic Divergence). A major flow of this upwelled water extends northward
as far as the Antarctic Convergence or Polar Front (Mann and Lazier 2006). The
Antarctic Convergence, which encircles Antarctica roughly 1,500 km off the coast,
divides the colder and fresher southern water masses and the warmer and saltier
northern waters; creating the largest pelagic boundary of the world ocean (Sournia
1994). Based on water mass properties within the Antarctic Circumpolar Current
(ACC) three major transitions are apparent, referred to as the Subantarctic Front,
the Polar Front, and the Southern ACC Boundary (Orsi et al. 1995). Changes in sea
surface height determined from satellite altimeter indicate each of these fronts is in
fact formed by three coherent fronts (Sokolov and Rintoul 2009). The most con-
spicuous open ocean fronts are those formed at the transitions between the pole-
ward extensions of warm-salty western boundary currents (e.g. the Gulf Stream;
Kuroshio; Agulhas and Brazil currents) and cold-less saline subpolar waters. As
the ACC deflects northward downstream of Drake Passage the Subantarctic Front
penetrates northward in the South Atlantic and nearly merges with the Subtropical
Front creating even more intense surface gradients. These fronts are characterized
by strong frontal jets and strong surface temperature, salinity and nutrient gradients,
and are frequently associated with intense eddies and meanders that developed by
instabilities of the mean flow (Fig. 1.1).

2.7 Frontal Eddies

Eddies and large-scale meanders are ubiquitous features of the ocean circula-
tion and naturally emerge from instabilities of the mean flow. There are several
classes of rings or eddies in the ocean, originated by different forcing and cov-
ering a range of spatial and temporal scales; we consider here just one type: the
frontal eddies. They contain pockets of moving water that break off from the main
body of a front and can travel independently, covering long distances before dis-
sipating. Eddies are commonly found in the vicinity of faster flowing currents
that form intense fronts with the surroundings waters, such as the Gulf Stream,
the Kuroshio Current, the Brazil Current, the Agulhas Current and the Antarctic
Circumpolar Current. Strong currents meander in a wave-like fashion and become
unstable; these flow instabilities lead to pinching off of relatively warm or cold
waters that act as a seed for frontal eddies. The water within such eddies has tem-
perature and salinity characteristics different from the surrounding waters. Frontal
eddies can take the shape of warm-core (masses of warm water turning within
colder ocean waters) or cold-core (masses of cold water within warmer waters)
eddies (Fig. 2.3c). Eddies nearly always contain embedded frontal interfaces, and
like other frontal types, embody mechanisms by which the physical energy of the
ocean system can be converted to trophic energy to support biological processes.
Recent high resolution observations and numerical models also indicate that rela-
tively short lived (~1 day) submesoscale structures (1–10 km) may significantly
12 2  Frontal types

contribute to pumping nutrients to the upper layer, subsequently increasing the


ocean primary productivity (Lévy et al. 2001, 2012; Klein and Lapeyre 2009;
D’Asaro et al. 2011). Fronts are therefore “hot spots” of intense biological and
physical activity (Olson and Backus 1985; Olson et al. 1994; McGillicuddy et al.
1988; Bakun 2006a).

2.8 Fronts Associated to Geomorphic Features

Some fronts are topographically controlled. When tidal or other currents interact
with irregularities in the sea bed, or coastline, it is usual to find consistent patterns
of eddies and associated fronts. These obstacles to the flow create accelerations
in the direction perpendicular to the upstream flow direction as well as frictional
boundary layers close to the obstacle. These forces, together with vertical strati-
fication and the Earth’s rotation determine the path of flow particles around the
obstacle. Thus, the presence of a headland, an island, a bank, a reef or an undersea
mountain cause a disturbance in the flow generating complex three-dimensional
secondary flows of various scales (Arístegui et al. 1994; Dong et al. 2007) that
result in a physical front (Fig. 2.3d) (Wolanski and Hamner 1988). Direct obser-
vations and numerical simulations indicate that these structures display enhanced
phytoplankton growth (Dong et al. 2009).
Chapter 3
Biology of Fronts

Abstract  Vertical movements that bring nutrient-rich waters into the well-lit sur-
face layers are at the base of the biological production of marine fronts; phyto-
plankton show strong positive reaction to such nutrient enrichment. The high
primary production generated is transferred then to higher trophic levels reach-
ing top predators, and also benthic organisms. High nutrient supply promotes the
growth of large-sized phytoplankton, and consequently the development of shorter
and more efficient food webs at fronts. Moreover, large-sized phytoplankton sinks
relatively fast, increasing the food supply to benthic assemblages. The largest and
more stable fronts are recognized as biogeographic boundaries, and those hav-
ing lesser spatial scales or persistence exert their influences at finer spatial scales
(ecoregions, assemblages). In most of the cases fronts do not appear to be abso-
lute barriers, but are leaky boundaries. The most accepted effects of fronts on
biodiversity are their impacts on divergences in species composition (β-diversity;
assemblages), while their effects on absolute measures of biodiversity seem to be
contradictory. Fronts result typically spawning grounds for species laying plank-
tonic eggs. They offer adequate conditions for the development of the early life
stages (abundant food; suitable physical-chemical ranges), and the possibility for
eggs and larvae to be retained near the front, both passively or by coupling verti-
cal migrating behavior to frontal circulation. Adult animals migrate to take advan-
tages of seasonal habitats; migrants could utilize fronts as marks or paths to guide
them in the highly dispersive and traceless pelagic realm. Animals may respond
to physical-chemical gradients and/or prey abundance to find their way along the
migration routes.

Keywords Nutrients vertical flux · Mixed layer turbulence · Organic matter


sinking  ·  Plankton concentration  ·  Size-structured trophic web  · Phytoplankton
size  ·  Food web length  ·  Herbivore food chain  ·  Microbial loop  ·  Pelagic bio-
geography  ·  β-diversity  ·  Bakun’s triad  ·  Plankton retention  ·  Migratory paths

© The Author(s) 2015 13


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_3
14 3  Biology of Fronts

3.1 Biological Production

Phytoplankton supports most of the trophic webs in the ocean (Smetacek 1999),
but major resources for those single-celled plants are heterogeneously distributed.
One of the problems confronting phytoplankton is their requirement of light and
nutrients for growth and reproduction, but the source of light is above, while the
source of nutrients is at depth (Margalef 1997). Phytoplankton inhabits the well-
lit upper few meters of the ocean, called the euphotic zone; the remainder of the
ocean is too dark to support net photosynthesis. Ambient light in the sea is loga-
rithmically attenuated by scattering and absorption of seawater which affect vari-
ous parts of the spectrum differentially with depth (Fig. 3.1). The euphotic zone
has a variable depth depending on several factors but it is always very thin relative
to the total depth of the ocean, extending hardly beyond 100 m in depth (Margalef
1997; Reynolds 2006). Turbulence within the surface layer moves phytoplank-
ton cells up and down, thus plants carried by vertical fluxes experiment different
light intensities while travelling at different depths. The depth to which plants can
be mixed and at which the total photosynthesis for the water column is equal to
the total respiration is known as the “critical depth” (Sverdrup 1953). Turbulence

Fig. 3.1  Light intensity decay exponentially with depth; consequently the illuminated zone
(euphotic layer) results in a thin stratum as compared to the total ocean depth. Organic matter
sinks in the oceans; most of remineralization occurs in deep and dark waters, modified from Lalli
and Parsons (1997)
3.1  Biological Production 15

modulates the light intensity experimented by phytoplankton cells. Consequently,


a phytoplankton population can only proliferate if mixing is shallower than such a
critical depth. On the other hand, particulate matter (the source for plant nutrients)
sinks in the sea (Fig. 3.1), and the export of organic matter from the upper ocean
to the deep ocean is an essential mechanism of the cycling of organic matter.
About 1–40 % of the photosynthetically fixed carbon sinks and is remineralized
in the deep ocean at substantially slower rates than in surface waters (Ducklow
et al. 2001). The resultant increase in dissolved inorganic carbon concentrations
towards the interior of the ocean, referred to as the biological pump, is regulated
by food web processes such as grazing. The tendency of particles to form aggre-
gates accelerates the downward movement and the flux increases gradually with
depth. The remineralization of sinking particles by heterotrophic bacteria occurs
mostly at depth, and enriches the deep water masses with nutrients essential for
phytoplankton (Eppley and Peterson 1979; Jackson and Checkley 2011; Tesi et al.
2012). Consequently, most parts of the ocean are characterized by very low con-
centrations of plant nutrients in the illuminated layer, and abundance of nutrients
in the dark deep levels (Margalef 1978) (Fig. 3.1). Moreover, if the density of phy-
toplankton cells exceeds that of the surrounding waters they will sink; and if sink-
ing is not countered by an upward current the cells will sink below the illuminated
zone. Hence the organization of the marine ecosystem is approximately defined
by gravity and light directions (Margalef 1997), and the entire living system is
crucially dependent on various processes by which organisms and materials are
transported and redistributed (Bakun 1996). Because marine fronts are character-
ized by relatively intense vertical circulations (Klein and Lapeyre 2009) and are
frequently associated with flow convergence, they provide mechanical energy that
contributes to the trophic energy balance of the biological community (Margalef
1978; Legendre et al. 1986; Bakun 1996).
Early observations of fronts emphasized the convergence of surface waters and
the associated downwelling, because any organism buoyant enough to resist the
downwelling would be expected to aggregate at the front (Fig. 3.2). Thus, it was
thought that passive advection was a possible explanation for the concentration of
planktonic organisms found at fronts. An alternative explanation invoked in situ
production, made possible by particularly favorable conditions of light and nutri-
ents (Mann and Lazier 2006). There are a number of processes (all linked to the
secondary1 circulation and hence not so easily distinguishable from each other) that
lead to the enrichment of nutrients for plants in the vicinity of fronts. The strong
horizontal pressure gradients generated at fronts are often balanced by the Coriolis
force, leading to strong along-front currents. When these two forces are not per-
fectly balanced, vertical circulations can be generated at the front. These vertical
circulations have the potential of carrying deep, nutrient-rich waters into the well-
lit surface layer, and stimulate phytoplankton growth (Sournia 1994). Both

1  Secondary circulation refers to the motion relative to a basic flow (geostrophic and hydrostatic
balanced).
16 3  Biology of Fronts

Fig. 3.2  Aggregation of planktonic organisms at fronts. The dark blue area indicates the region
with the highest abundances. f front; arrows indicate currents and dashed lines are lines of equal
density

processes (aggregation or in situ production) can simultaneously occur at the same


front. Moreover, frontal instabilities can lead to vertical restratification, which, if
overcome turbulent mixing, can trigger frontal blooms (Taylor and Ferrari 2011).
Fronts are zones of increased lateral and vertical mixing, usually leading to
increased primary and secondary production (Olson and Backus 1985). It is well
accepted that fronts are characterized by high phytoplankton biomass and in most
cases, also enhanced activity at higher trophic levels (Le Févre 1986; Largier
1993; Acha et al. 2004). Nutrient rich waters are upwelled in fronts; and if the
frontal region is sufficiently long-lived, populations of herbivorous zooplankton
will increase, and convergence will concentrate zooplankton in the front promot-
ing secondary production. Unique biological properties of fronts have for many
years been associated with red tide outbreaks, typically concerning dynoflagel-
lates, and mostly in coastal waters (Pingree et al. 1975; Le Févre 1986; Sournia
1994). Along front fluxes cause the advection of low trophic level organisms (i.e.
phyto- and zooplankton). Maintenance of a population in a unidirectional large-
scale flow demands population dynamics that utilize counter-flows and the struc-
ture of the eddy field to provide a steady seed population or recruitment to the
upstream end of the domain. The dynamic tendency for the large scale flow in
frontal zones to become unstable (i.e. breaking down into meanders and eddies),
provides an appropriate physical setting that allows this recruitment; then dissimi-
lar responses by phyto- and zooplankton to the flow field generate different abun-
dance patterns along the front (Olson et al. 1994).
3.1  Biological Production 17

In late winter in mid to high latitudes, when phytoplankton growth is limited


by lack of sufficient light intensity the upper water column tends to be well mixed.
Under these conditions the onset of stratification leading to the spring bloom is
thought to be triggered by net heat flux through the sea surface. The lack of verti-
cal stratification required to support phytoplankton close to the illuminated sur-
face layers, however, can also be overcome by lateral mixing across fronts. Recent
observations from the subpolar North Atlantic and biophysical models indicate
that the initial stratification and resulting bloom can be caused by eddy-driven
slumping of the cross-front density gradient (Mahadevan et al. 2012). These analy-
ses show that frontal eddies exchange cold-dense water equatorward and warm-
light water poleward creating shallow mixed-layers about 20–30 days earlier than
would occur by surface warming. The combination of very weak vertical stratifi-
cation and turbulent regime that characterizes subpolar waters limits the growth
of phytoplankton regardless of the relatively high nutrient concentrations, creat-
ing the so-called high-nutrient low-chlorophyll (HNLC) environments. At the
transition between oligotrophic subtropical and HNLC subpolar waters, however,
small-scale cross-frontal mixing can create layers with sufficient nutrients and
stratification to promote phytoplankton growth (Brandini et al. 2000).
Secondary circulations in frontal areas not only promote fertilization by nutri-
ents, but also may result in downward export of particles and organisms toward
subsurface layers and account for the persistence of large populations of both
invertebrate and vertebrate species at depth (Sournia 1994). Pelagic productivity
and physical processes largely determine the quantity and quality of organic mat-
ter reaching the seafloor, which can be derived from a variety of sources, includ-
ing phyto- and zooplankton remains, crustacean molts, macro-aggregates (‘marine
snow’), and fecal pellets; all originating in the primary productivity by phyto-
plankton (Berkenbusch et al. 2011). In general terms, the flux of organic carbon
is directly linked to surface-water productivity and the supply of organic matter
generally decreases with increasing water depth. The importance of fronts for ben-
thic communities arises not only from the high primary production but is also due
to the vertical fluxes that transport food particles to the seafloor, and due to the
weakening of the vertical stratification that increases the sinking rates of particu-
late matter. Once on the bottom, the relatively un-degraded material is rapidly con-
sumed and incorporated into benthic biomass.
Enhanced biological production occurs at fronts as a consequence of the match-
ing or resonance of physical scales with biological scales (Legendre et al. 1986).
The length of the organism’s generation relative to the persistence and predictability
of a front will determine whether its response is by population growth or behavio-
ral mechanisms (Angel 1986). Plankton generation time is of the order of days or
weeks (Legendre et al. 1986) consequently their abundances at fronts is partly due to
population growth (in some fronts concentration by convergence is also important).
Medium and larger predators actively seek fronts, showing a behavioral response.
High food availability at fronts attract nekton organisms (e.g. fish, squids) transfer-
ring the energy to higher trophic levels. The response of organisms to fronts inte-
grates simultaneous reactions to main physical, chemical and biological gradients,
18 3  Biology of Fronts

thus relationships between organisms’ distributions and fronts may be complex and
nonlinear, especially for more mobile species (Brandt 1993).
Free swimmers like tunas, swordfish or sperm whales, detect fronts by sophis-
ticated sensorial systems (Olson 2002). Strong convergence velocities associated
with fronts are very efficient in accumulating not only plankton but also other
floating materials along the convergence line. Flotsam often includes detritus
such as dust, foam and timber (Bowman 1978). Fish and other marine animals
show widespread attraction to drifting objects (Fig. 3.3), and this could be used
as a front detection mechanism in some cases (Bakun 1996; Castro et al. 2002);
so fish aggregated to drifting objects may obtain food by preying on organisms
aggregated at fronts. Other pelagic species, generally the long range migrants such
as tuna, may use floating objects as a landmarks or “meeting points” to increase
the encounter rate between isolated individuals or small schools and other schools
of con-specifics, thereby forming large schools to continue upon their migration
routes (Castro et al. 2002). However, the roles played by environmental variables
and by behavioral processes (e.g. social behavior) in the formation of these aggre-
gations remain elusive (Robert et al. 2013).
Megaplanktivores such as filter-feeding sharks, Manta rays and baleen whales
are at the apex of a short food chain (phytoplankton–zooplankton–vertebrate) and
are sensitive indicators of sea-surface plankton availability (Fig. 3.4). Finding suf-
ficiently large concentrations of appropriate prey in the open ocean to meet their
high metabolic needs is an impressive skill. It has been demonstrated that predicta-
ble oceanic and inner-shelf fronts are principal feeding areas for those megaplank-
tivorous species (Sims et al. 2005; Bost et al. 2009; Graham et al. 2012). Elephant
seals may use frontal eddies as foraging areas (Campagna et al. 2006; Bost et al.
2009). Marine turtles can exploit fronts as forage habitats (Polovina et al. 2001;

Fig. 3.3  Fishes attracted around old fishing rope (Simon Max Bannister and Sara Close, 5 Gyres
South Atlantic expedition). Strong convergence velocities associated with fronts are very efficient
in accumulating floating materials. Fishes show widespread attraction to drifting objects, and this
could be used as a front detection mechanism in some cases
3.1  Biological Production 19

Fig. 3.4  Megaplanktivores are at the apex of a short food chain and are sensitive indicators of sea-
surface plankton availability. a Manta ray (Manta birostris) (Guy Stevens http://www.mantatrust.
org/about-mantas/feeding-frenzy/). b Whale shark (Rhincodon typus) (Aquarium of the Pacific http:
//www.aquariumofpacific.org/onlinelearningcenter/species/whale_shark). c Basking shark (Cetorhi-
nus maximus) (Prionace.it http://www.prionace.it/squaloelefanteENG.htm). d Humpback whales
(Megaptera novaengliae) (NOAA http://sanctuaries.noaa.gov/jointplan/presskit/welcome.html)

Ferraroli et al. 2004). The sea snakes Pelamis platurus concentrate in areas of
frontal convergences, reaching hundreds or even thousands of animals in surface
slicks. Convergent flows concentrate also floating debris that in turn attracts small
fishes which are preyed upon by the snakes (Dunson and Ehlert 1971). Diving
birds like penguins also use fronts, including frontal eddies, as forage grounds
(Bost et al. 2009). Coastal birds like gulls and terns, may detect prey aggrega-
tions by using visual cues, both by direct location of prey or identifying the active
presence of other subsurface predators (large fish, seals, whales, dolphins and
even penguins) which drive preys close to the surface. Pelagic seabirds, like large
albatrosses and petrels, may travel thousands of kilometers during their foraging
flights and may use olfactory cues to detect remote sources of food like zooplank-
ton, fishes or squids which concentrate at fronts (Nevitt 1999). The occurrence of
productive and predictable fronts may also promote the establishment of breeding
colonies in their proximities (Russell 1999; Bost et al. 2009).
Some similarities and differences between terrestrial and marine systems may
be established that can help understanding of properties and importance of marine
fronts. In a forest, most of the life concentrates at the uppermost layer, because
there is where most light is available, and every organism living below depends
20 3  Biology of Fronts

on this layer’s production. Thus, the canopy is comparable to the euphotic layer
in the sea, while below lies an equivalent of benthos, where organisms live
entirely on secondary materials that fall from above (or roots and fungi). The
basis of such analogy is the three-dimensional character of the forest and the
ocean, and the equivalent significance of the vertical dimension in both types of
habitats (Fig. 3.5). In both cases, the maximum thickness in which the light can
be used and the chlorophyll is active is about 100 m high (Bates 1960; Margalef
1997). The main difference between primary producers of both types of ecosys-
tems lies in the microscopic size of marine phytoplankton organisms, which
contrasts with the much larger size of terrestrial plants. The biomass of the trees
comprise a major fraction of transport and supporting tissues, which forms part
of a very effective system that brings nutrients from the ground to the lit can-
opy. Phytoplankton, on the other hand, appears as a community which is poorly

Fig. 3.5  Both in the forest and in the sea photosynthesis occurs in the well illuminated upper
part (the canopy and the euphotic zone, respectively) while remineralization takes place in the
lower and shadowy stratum. A main difference is that in terrestrial ecosystems the upward nutri-
ent transport is internal and under plant’s control, while phytoplankton depends on the mechani-
cal energy of the ocean for bring nutrients in the euphotic zone
3.1  Biological Production 21

controlled from inside. Plankton contains nothing comparable with the well-
structured transport system found in a forest. In trees, the major part of power that
brings up water and mineral nutrients to the leaves comes from the pull caused by
evaporation of water at the leaves; evaporation here plays an equivalent role to the
work done by turbulent energy or upwellings in aquatic environments (Margalef
1997). In natural plankton, control of nutrients transport is still entirely in the
physical environment, in the mobile structure of water masses, with cells of circu-
lation and eddies of every size. In terrestrial plants, the transport system internal-
izes the nutrient cycle and places it under the plant’s control, unlike what occurs
in the sea (Margalef 1978, 1997) (Fig. 3.5). This underlines the ecological impor-
tance of marine fronts, characterized by upward water movements that fertilize the
illuminated zone.

3.2 Trophic Webs

While recognizing that there are some highly specialized marine predators, the
diverse diet of many species indicates that feeding at sea is often opportunis-
tic and can be considered as less dependent on prey taxonomy than on prey size.
Contrasting with changes in species composition, the size spectra of marine eco-
systems exhibit remarkably constant shapes. This observation suggests that,
beyond strict species interactions, size-based interaction controls the energy trans-
fer in the marine environment (Cury et al. 2001).
With some notable exceptions (Sargassum for instance), most of the primary
organic production in the open sea is by single-celled plants and most herbivores
are small, but they are usually larger than the plants (Sheldon et al. 1977). In oli-
gotrophic, oceanic waters the base of the food chain is composed of very small
cells. Smallness is usually seen as an adaptation to take up extremely low nutri-
ent concentrations because of a favorable surface to volume ratio. Such kind of
phytoplankton is too small to be ingested by copepods thus most of primary pro-
duction is channeled through the “microbial loop” (picoplankton-heterotrophic
nanoflagellates-ciliates). In these environments, copepods feed mainly on hetero-
trophic nanoflagellates and ciliates, thus adding links to the classical food chain
and raising the trophic level of zooplanktivorous fish. Whenever there is an input
of nutrients, such as in fronts, larger phytoplankters (i.e. diatoms; dynoflagellates)
become dominant. Most of such phytoplankton falls well into the food spectrum
of herbivorous copepods, and consequently those “classical” food chains consist
of fewer levels (Fenchel 1988; Kiorboe 1993; Sommer et al. 2002).
The size structure of phytoplankton depends not only on nutrients abundance
but also on hydrodynamic forces. Sinking of organic particles out of the euphotic
zone represents a major loss of organic matter to the deep ocean and eventually to
the sea floor; therefore, large cells depend on water motion to remain suspended.
Both empirical observations (Taylor et al. 2012) and models (Rodriguez et al. 2001)
indicate that the relative proportion of large phytoplankton cells increases with the
magnitude of the upward velocity. This suggests that mesoscale vertical motion
22 3  Biology of Fronts

(a ubiquitous feature of fronts) may aid in controlling the size structure of phyto-
plankton near fronts affecting the structure of the food web (Rodriguez et al. 2001).
Fronts stimulate the growth of large-sized phytoplankton, mainly composed by
fast growing bloom specialists (Dutkiewicz et al. 2009), and support a classical food
chain with the dominance of herbivorous forms like copepods, euphausiids or appen-
dicularians. Away from fronts, there is a tendency for the dominance of mostly car-
nivorous or omnivorous taxa in the zooplankton assemblages (Ohman et al. 2012).
Thus, vertical flows at fronts provide nutrients and turbulent energy that promote not
only high primary production but also shorter and more efficient food webs in which
a larger proportion of the primary production is channeled to larger organisms.
In most fronts the generated biomass is exploited by a trophic web involving,
at the higher trophic levels, highly mobile organisms (e.g. fish; squids; birds), and
as a consequence of their migrations the biomass produced at the front is exported
to remote oligotrophic areas. In this way, biomass over-accumulation is avoided
in the front and the exploited community remains juvenile and opportunistic.
The interaction of two parts of the system, a productive one and a consumer and
mobile one, may be considered as an exploitation (=biomass exportation) of an
ecosystem by another one. This is why the total effects of biological production at
fronts are hard to estimate, and ecological phenomena occurring at fronts may be
more important in determining the ecological properties of the area than the phe-
nomena occurring inside the two adjacent water masses (Frontier 1986).
Phytoplankton size can also affect food supply to the benthos. Large cells sink
faster than small ones, since sinking rates increase exponentially with the cell size
and because larger cells aggregate more rapidly into sinking flocs than small cells.
Consequently, a greater fraction of the large cells primary production may sink out
of the water column (Kiorboe 1993). The production of large cells (e.g. diatoms)
at fronts could enhance the pelagic-benthic coupling because phytoplankton cells
may drift away from the upwelling core, promoting the arrival of phytoplankton to
the sea bed and increasing the heterotrophic activity of benthic communities.
It is virtually universal that among the plankton and the nekton a predator is
larger than its prey, although this is not so generally true of the benthos. Marine
animals live in a medium that is eight hundred times denser than air, where only
a streamlined morphology allows active and efficient movements (though small
plankton living under low Reynolds numbers are an exception to this (Mann and
Lazier 2006); this is why the development of appendages to handle and capture
large-sized prey is not common. Thus, a pelagic predator must have a jaw large
enough to swallow its prey as a whole. As the size of the jaw is related to the organ-
ism’s size, the predation process is believed to be largely determined by the size
ratio between predators and prey (Sheldon et al. 1977). On the other hand, feeding
on too small sized preys could be inefficient unless predators have special struc-
tures to concentrate small preys (e.g. the gill rakers of filter feeding fishes such as
sardines and anchovies; or the baleens of true whales), and at most sites predators
select large preys that they can swallow, even when small ones might be present.
Marine fronts may affect predator distribution by augmenting the profitability
of small sized preys. Many predators are threshold foragers, which require prey to
3.2  Trophic Webs 23

exceed a minimum density before they can forage successfully. In the neighbor-
hood of fronts predators may feed upon small prey present in high density because
frontal dynamics aggregate food particles near the surface; thus, fronts may create
new foraging opportunities by concentrating prey items that are usually too small
to be profitably fed upon if concentrations are low (Vlietstra et al. 2005). Although
predators may prefer relatively large prey items over small ones, the amount of
energy per unit volume of water of small prey may equal or exceed that of large
prey when marine fronts cause small prey to concentrate in dense patches. Fronts
may therefore represent profitable foraging sites for some predators (Vlietstra
et al. 2005), turning otherwise unprofitable sections of the marine landscape into
profitable foraging grounds.
Different fronts are able to concentrate different prey types and this deter-
mines their use by predators of diverse body size and diet. In the Southern Ocean,
large marine birds such as albatrosses and gadfly petrels, which partially depend
on squid, dominate seabird biomass near the Subtropical Front; while small spe-
cies that feed primarily on macrozooplankton, such as prions, dominate near the
Polar Front. The highest concentrations of blue whales (Balaenoptera musculus),
humpback whales (Megaptera novaeangliae), Fin whales (B. physalus) and Minke
whales (B. bonarensis) have been found in strong association with the Antarctic
Divergence, reflecting the distribution patterns of their main prey, the Antarctic
krill. Sei whales (B. borealis) on the other hand, having a diet dominated by cope-
pods, are concentrated close to the Subtropical Front (Bost et al. 2009).
The food web is a composite not only of trophic levels but also of organisms
with differing time scales of life cycles; thus predator/prey interactions imply
interactions across scales (Steele 1989). Variations in the timing of events play an
important role in marine life, including the rhythms of nutrient enrichment pro-
cesses in fronts that can affect trophic webs. Le Févre and Frontier (1988) con-
sidered a tidal front in which the fertilizing mixing process occurs in a 14 days
cycle, matching the neap-spring tides, and a shelf-break front, where the fertiliza-
tion process is of high frequency, 12 h periodicity matching the semi-diurnal tides.
Based on distributions of zooplankton biomass they concluded that in the latter
case enhanced productivity was in the form of a classical herbivore food chain,
while in the former case primary production was consumed by microorganisms,
because herbivorous copepods cannot adapt to short-lived, fortnightly phytoplank-
ton blooms.

3.3 Biogeography

There are several differences in the distribution patterns of terrestrial and marine
organisms. These differences are not surprising since distinct mechanisms are at
work in these two realms. On land, critical habitat characteristics can change drasti-
cally over short distances because of explicit barriers such as mountains, deserts,
and watergaps. In other cases, the clustering of terrestrial range limits may derive
24 3  Biology of Fronts

from historical events such as glacial intrusion or land bridge submergence. In


marine systems, however, it becomes more difficult to envision how persistent range
boundaries can become locally concentrated. Although substrate types vary spatially
and rivers locally alter salinity, a single, continuous, dispersal medium (the ocean)
connects all available habitats, and environmental gradients within this medium are
neither as contrasting nor as immutable as on land (Gaylord and Gaines 2000).
The idea that fronts separate different water masses, hence different pelagic
populations, is an old one and it is sustained by a number of observations at vari-
ous scales (Sournia 1994). However, the role of fronts in marine biogeography is
not yet fully understood. Conceptually, oceanographic mechanisms acting across a
range of spatial and temporal scales can be viewed as a base level of environmen-
tal structure that influence a suite of biological processes that in turn influence the
formation of macroecological patterns, both directly and indirectly (Leichter and
Witman 2009). Though the tight linkages between biogeography and a small num-
ber of ocean drivers is recognized (Spalding et al. 2012), there is a lack of compre-
hensive interpretation of the function of fronts in setting biogeographic boundaries.
Longhurst (1998) pointed out that subdivisions based on oceanographic criteria (pri-
marily the positions of fronts) may be appropriate in certain circumstances. Some
fronts seem to constitute biogeographic boundaries such as the Antarctic Polar Front
(Sournia 1994; Boltovskoy et al. 2005; Bost et al. 2009; Spalding et al. 2012), or the
shelf-break fronts that usually mark the neritic-oceanic transition (Longhurst 1998;
Spalding et al. 2007) however fronts having lesser spatial scale or persistence do not
constitute the boundaries between the geographic units in the most recent biogeo-
graphic marine systems (Longhurst 1998; Spalding et al. 2007, 2012). This means
that in some way organisms are able to cross fronts. Notwithstanding, fronts have
been recognized in finer resolution regional classification systems such as ecore-
gions (Spalding et al. 2007) or faunal assemblages’ boundaries. For example, zoo-
plankton assemblages including chaetognaths, salps, krill larvae and copepods may
be strongly influenced by frontal structures (McGinty et al. 2011). Frontal disconti-
nuities may also manifest in terms of the developmental stages, as seen in the distri-
butions of eggs, larvae and juveniles of fishes across the Ushant front or by different
copepodite stages in the Ligurian Sea (Sournia 1994).
Most efforts have been devoted to study the role of fronts on spatial patterns of
plankton (Hunt and Hosie 2003; Boltovskoy et al. 2005), but frontal effects have
been also reported for pelagic fishes (Moteki et al. 2011), demersal fishes (Gaertner
et al. 2005), benthic invertebrates (Boltovskoy et al. 2005) and marine birds (Piatt and
Springer 2003; Bost et al. 2009). Physical contrast across a front may determine its
influence on the community structure. Minor physical differences that cannot generate
changes in the specific composition of the community may produce major changes
in the dominance pattern across fronts (Angel 1986). The degree to which different
groups of organisms are affected by fronts is also variable. Antartic endemism for
benthic invertebrates is higher than for zooplankton, showing that the effects of the
Antartic Polar Front are different for such groups (Boltovskoy et al. 2005).
In most cases fronts represent a leaky boundary between different ecological
regimes and several data indicate that fronts exert mixing effects on the adjacent
3.3 Biogeography 25

populations. At first it may seem rather counter-intuitive that fronts are where
waters mix, but several studies have demonstrated the extent to which parcels of
water, containing biota, pass from one side to other by a variety of mechanisms
(Ashjian 1993; Sournia 1994; Longhurst 1998). For example, despite the impor-
tance of the Gulf Stream as a biogeographical boundary, it appears to function as
a leaky interface permitting some cross-stream exchange of water and therefore,
plankton populations. The meanders of the Gulf Stream may be sites of cross-front
exchange of plankton populations between the Sargasso Sea and the Slope Water
(Ashjian 1993). Another proposed mechanism is that instead of a slow continuous
exchange, many fronts accumulate material for a while and then through an evolv-
ing instability lead to a single large pulse of cross-front exchange. This pulsed
exchange may be as effective in bringing about exchange across the front as if
there were no front (Largier 1993). Though it is commonly observed that different
groups of species and different ecological conditions dominate on either side of
fronts, it is also commonplace to observe that individuals of many, perhaps most,
of the relevant species can also be found in small numbers on the opposite side of
the front. Given the dynamic exchange of water across fronts at all scales, it could
hardly be otherwise (Longhurst 1998).
Though there are numerous studies about fronts and the spatial patterns at dif-
ferent scales of planktonic, benthic and nektonic organisms, such information has
not been yet systematized nor analyzed in a comparative way. This could be in part
the reason why mechanistic explanations on how fronts can create or influence those
patterns are limited. It is commonly assumed that gradients in environmental condi-
tions are the primary determinant of the boundaries of the ranges of species, particu-
larly when species’ boundaries cluster at a given location. In setting such boundaries
there are two classes of causes: one based on mortality outside the specie’s range,
either due to physical, chemical or biological processes, the other based on barriers
to larval dispersal. Unfortunately, the underlying oceanographic mechanisms poten-
tially responsible for these two causes of range limits—steep physical gradients
versus hydrographic barriers to dispersal—are typically confounded in space. Steep
gradients in ocean temperature or other physical parameter cannot be generated and
maintained without anomalous circulation patterns (e.g. convergent currents) which
tend to restrict larval dispersal (Gaines et al. 2009). Another point of view is that
distributional patterns of populations of marine species with complex life histories
(i.e. those with planktonic egg and larval stages) are controlled by oceanographic
processes that facilitate birth site fidelity to reproductive grounds (Sinclair 1988).
Many fishes and invertebrates having planktonic larvae, which represent most of
the cases, choose fronts as spawning grounds (see the Section on Larvae retention).
Thus, marine fronts may play a role in setting populations’ spatial structures but not
necessarily being or defining the borders of the species’ geographical distributions.
Although the current systems capture the major elements of the biogeographic
patterns (i.e. provinces), considerable further “texture” does exist in the oceans
at smaller scales, including fronts (Spalding et al. 2012). It seems that the role of
fronts in setting biogeographic boundaries depends on their spatial scale, physical
contrast, and persistence. As those properties increase, so does the frontal influence.
26 3  Biology of Fronts

3.4 Diversity

Studies focused on the effects of marine fronts on diversity are scarce, and their con-
clusions are somewhat contradictory. Most reports indicate that fronts play a role
in setting up diversity patterns, while other studies suggest that these patterns occur
on large scales not necessarily associated with typical cross-front scales (Stemmann
et al. 2008; Mauna et al. 2011). Several studies ascribe diversity patterns to front
occurrence; involving different groups such as phytoplankton (Ortner et al. 1979;
Jeffrey and Hallegraeff 1980); zooplankton (Ortner et al. 1978; Tranter et al. 1983;
Gaard et al. 2008; Hosia et al. 2008); hyperbenthos (Dewicke et al. 2002); fish larvae
(John et al. 2001; Sánchez-Velasco et al. 2012); cephalopods (Brandt 1983); rays and
sharks (Lucifora et al. 2012); midwater fishes (Olson and Backus 1985); demersal
fishes (Alemany et al. 2009); tunas and billfishes (Worm et al. 2003); and seabirds
(Haney 1986). These findings refer to several types of fronts, and the diversity pat-
terns were expressed as divergences in species composition (e.g. β-diversity or dif-
ferent assemblages); or as diversity in absolute terms (measured as α-diversity (e.g.
species richness), or in defining hotspots). High mean species richness and diversity
of whales and seabirds are consistently associated with fronts at the Southern Ocean
(Bost et al. 2009). In the case of predators, high diversity at fronts is in general attrib-
uted to the high biological production and better feeding opportunities because an
abundant prey supply acts as an attractor to individual species and, at the same time,
allows for the coexistence of a high number of predator species (Lucifora et al. 2012).
Fronts concentrate high biological activity but from a conceptual point of view,
this does not necessarily imply that they show higher diversity of species as com-
pared with neighboring environments. Several forms for the relationship between
species richness and productivity have been proposed, but none are generally
accepted. A greater variety of species may be expected in very productive envi-
ronments because more resources can allow more species to coexist (Wright et al.
1993). However, hump-shaped patterns have also been described whereby as pro-
ductivity rises, diversity first increases and then declines (Rosenzweig and Abramsky
1993; Gaston 2000) due to increased competitive exclusion (Abrams 1995).
Therefore, high ecosystem productivity can lead to either an increase or decrease in
species richness, or a combination of both (e.g. a hump-shaped distribution).

3.5 Life Histories Traits in Relation to Fronts

A major problem in evolutionary biology is to explain the amount and structure of


biodiversity in widely connected environments like the open ocean. The effects of
fronts on certain organisms could partially explain specific and population diver-
sity. Even though in most cases fronts do not appear to be an absolute barrier to
pelagic organisms present on either side (Sournia 1994), strong temperature gra-
dients across fronts could act to uncouple life-cycle events, including reproduction
(Gaard 1996). In planktonic species whose range spans a strong hydrographic front,
3.5  Life Histories Traits in Relation to Fronts 27

the timing of reproduction may be offset sufficiently on both sides of the front to
produce effective genetic isolation even in the face of continued dispersal. It is also
conceivable that a species could experience sufficiently strong selection on either
side of a hydrographic boundary to produce different body size, growth rates, or
skeletal shapes with consequent changes in mating recognition systems, particu-
larly because these morphologic variables are known to be affected by changes in
food supply, temperature, and predation intensity. Hence it is possible that pelagic
speciation occurs in the face of sustained gene flow that is rendered ineffective by
changes in mating recognition cues or reproductive timing (Norris 2000).
Regarding highly mobile, nektonic organisms (e.g. fishes, squids), species in
general inhabit extended geographic regions. Populations undergo migrations, sea-
sonally vacating and reoccupying specific sub regions of their respective ranges.
Marine fishes and other oviparous organisms select the environment in which their
eggs will be released (Roosenburg 1996), so most organisms have a tendency to
undertake extensive movements to specific breeding sites (Breder and Rosen
1966). Spawning locations and subsequent larval distributions are associated with
well-defined and geographically predictable or stable oceanographic systems
(Sinclair 1988). This probably occurs because areas suitable for adult feeding may
not necessarily be suitable for the survival of early stages (Bakun 1996).
Marine fronts have been broadly reported as preferred spawning grounds for
fishes and squids (Sinclair 1988; Bakun 1996, 2006b; Acha et al. 2004; Houde
2009). Most fronts seem to fulfill the requirements of the “fundamental triad
hypothesis” that identify suitable spawning habitats (Bakun 1996): (i) nutri-
ent enrichment processes, (ii) concentration of food particles, and (iii) retention
of eggs and larvae within a favorable habitat. Although fronts are diverse in spa-
tial and temporal scales, and driven by varied forcing, recurrent features of these
scenarios are: (i) nutrient pumping due to stratification weakening or disruption,
generating enrichment in the euphotic zone that enhances primary production, (ii)
convergence of water masses that aids in concentration and maintenance of food
particles for larvae, and (iii) the existence of a vertically structured dynamics that
allows for behaviorally mediated larvae retention (Largier 1993; Mann and Lazier
2006; Bakun 2006b). As such fronts can include the whole “triad”.
Since most marine animals have a pelagic larval stage, the paradigm until recently
has been to assume extensive dispersal and massive export. In combination, the wide-
spread existence of planktonic larvae, the broad distribution of larvae in the plank-
ton, extended planktonic periods and poor swimming abilities of most larvae suggest
that the larval exchange among populations should be the rule. Consequently, the
concept of “open populations”, with plentiful exchange of larvae, was pervasive in
the late twentieth century. However, evidences from a variety of fields indicated that
local retention may be considerably more prevalent than previously thought, even in
species with long larval durations and, thus, that populations may be less open than
originally thought (Warner and Cowen 2002; Levin 2006). This is in agreement with
recent results based on molecular phylogenetic analyses that have revealed high cryp-
tic biodiversity in the open ocean, and that rates of speciation can also be as high for
pelagic taxa as for shallow-marine and terrestrial species (Norris 2000).
28 3  Biology of Fronts

The most striking difference between aquatic and terrestrial mating is that
aquatic organisms commonly shed female gametes as well as male gametes.
Though water is a benign medium for gametes, external fertilization is by no
means easy, especially for sedentary or completely sessile organisms (Strathmann
1990). Fertilization by distant males and females is limited by the life span of
active sperm, predation on gametes, and dispersion of gametes with dispersion
probably the greatest obstacle: classical diffusion principles imply that particu-
lates (e.g. gametes) should disperse widely over a vast area, somewhat like a cloud
of ever-increasing dimension (Wolanski and Hamner 1988). All these problems
diminish with reduced distance among individuals (i.e. increasing organisms’ spa-
tial density). It can therefore be argued that for free-spawners the high fertiliza-
tion success in crowded populations offsets the reduced fecundity from resource
limitation due to increased competition (Strathmann 1990). For benthic animals,
there is adaptive advantage in settling out from the plankton in optimal conditions
for adulthood. Hence, mechanisms permitting some settlement near to a sustained
parent population might reasonably also be expected to have been selected during
evolution (Naylor 2006). Moreover, if reduced larval dispersal resulted in reduced
genetic exchange among populations, it could increase possibilities for local adap-
tation (Strathmann 1990). Though settlement near parent population could increase
intra-specific competition; benefits from retention could surpass its disadvantages.
Because the ocean is a highly dispersive environment, drifting and mobile
organisms are continuously dispersed with the consequence that the distance to the
nearest mate persistently increases; thus the chance that one individual encounters
another with similar genetic material decreases monotonically as time after birth
increases. Diffusion itself from a point source for nonmobile drifting organisms,
or random movement for mobile organisms, minimizes the frequency of sexual
encounter that is necessary to allow persistence of the population. Survival itself
is not the only issue; finding a mate in a diffuse environment at low concentrations
becomes the additional, perhaps more critical, challenge. Thus the very existence
of a population may depend on the ability of larvae to remain aggregated during
the first few weeks/months of life (Sinclair 1988).
Water dynamics of marine fronts offer opportunities for planktonic organisms
(including larvae of fish and benthic animals) to be retained. There are frequently
steep gradients in flow velocity and even reversals in flow direction associated with
fronts (McManus and Woodson 2012). Plankton is a highly diverse group whose
components display a wide range of behavioral capabilities that bridge the transi-
tion from being a passive particle to being able to determine vertical and horizontal
position in the ocean (McManus and Woodson 2012). If an organism is not a pas-
sive particle (i.e. has the ability to float, sink or swim) then the potential exists for
the organism to become concentrated in certain types of flow (Franks 1992). Tiny,
weakly swimming organisms that may be unable to resist being passively swept
along in the horizontal ocean flow may well be able to control their depth level in
the much less energetic field of vertical motion in the ocean. Estimates of verti-
cal velocities at fronts from field studies and modeling indicate maximum speeds
of ca. 0.2 mm s−1. This is the same order of magnitude as the swimming speed
3.5  Life Histories Traits in Relation to Fronts 29

of most ciliates, flagellates, coccolithophorids, and diatoms (sinking). Most crusta-


cean (e.g. copepods, euphausiids, amphipods) as well as many types of fecal pel-
lets (sinking) have vertical velocities 1 or 2 orders of magnitude higher than the
vertical velocities at fronts (Franks 1992). Such organisms will thus accumulate
in the slowly sinking waters of the convergent frontal interfaces (Fig. 3.2) (Bakun
1996; Olson 2002; McManus and Woodson 2012). Larger organisms can break
through vertical density gradients (pycnoclines) between different water masses;
thus vertical migrations between water masses moving in different directions at
tidal, diel or longer timescales permit retention of planktonic larvae and adults in
favorable ecological locations (Fig. 3.6). Exogenous factors serving as cues for, or
directly controlling, vertical migrations rhythms include light, hydrostatic pressure,
salinity, temperature and water movements. The interaction of vertical migrating
behavior with vertically structured ocean transport processes offers a mean by
which living organisms are potentially able to follow drift trajectories that may in
no way resemble those followed by passive particles (Sinclair 1988; Bakun 1996;
Naylor 2006). Flexibility of behavior in response to hydrographic conditions
gives larvae unexpected freedom from normal restraints in controlling their move-
ments. Models that incorporate vertical migration often show that vertical move-
ments have a significant effect on plankton transport, and can lead to retention
which would not otherwise occur (Levin 2006). Frontal zones are characterized
by complex internal structure and may incorporate features across a wide range of
spatial scales. For example, the Mississippi River plume front consists of a large-
scale (2–20 km width) frontal zone within which small-scale (10–50 m wide) and
ephemeral convergence zones (Govoni and Grimes 1992) are embedded. Densities
of larval fishes within the large-scale frontal zone are probably the result of their
accumulation along ephemeral convergence zones and subsequent dispersal and
mixing during relaxation of convergence, so the spatial distribution of larvae in the
vicinity of the front is the aggregate result of the repeated formation and relaxation
of small-scale convergence zones (Govoni and Grimes 1992). Olson et al. (1994)
also distinguished between large-scale frontal zones and fronts, arguing that the

Fig. 3.6  In a counter-current
system, simple behavioral
traits (e.g. vertical migrations
associated to the day/night
passage) can generate
plankton retention, modified
from Weinstein et al. (1980)
30 3  Biology of Fronts

primary biological response is tied to the dynamics of the smaller scale features,
that is  the individual fronts, which are characteristic of these frontal zones.
Persistent fronts likely set recruitment patterns of those organisms having a
planktonic larval phase through a variety of biophysical coupling mechanisms.
As pointed out above, fronts can act as accumulators of passive, buoyant parti-
cles and weakly swimming organisms, in particular phytoplankton and small zoo-
plankton including larvae. Increased phytoplankton and zooplankton biomass at
frontal regions may also lead to increased development rates and shorter pelagic
larval duration for many species. Because recruitment of larvae to adults’ popu-
lations is a key component of resilience in marine ecosystems, regions of high
frontal activity, where recruitment is less variable and generally higher, will be
more resilient and sustainable compared to regions with low front probability
(Woodson et al. 2012).
The distributional problems faced by plankton are an important component of
the ecology of the oceans, not only because the plankton is a diverse and highly
abundant group occupying key links in trophic webs, but also because most of the
benthic or nektonic forms have planktonic larvae. Retention areas may be impor-
tant for maintaining population persistence, and the retention of eggs and larvae
in favorable areas has been hypothesized as an important determinant of marine
fish year class strength (Bernatchez and Martin 1996; Iles and Sinclair 1982).
Moreover, density relations can be particularly important and concentration of
organisms in frontal areas promotes biological interactions such as predation
(McManus and Woodson 2012); reproduction (Sinclair 1988) or parasites trans-
missions (Díaz Briz et al. 2012). As was nicely stated by Kiorboe (2008) “Life
is all about encounters… phytoplankton cells need to encounter molecules of
nutrient salts and inorganic carbon; bacteria need to encounter organic molecules;
viruses need to encounter their hosts; predators need to encounter their prey; and
males need to encounter females (or vice versa)”. Concentration and retention
processes (together with nutrients fertilization) are reasons for the key ecological
importance of marine fronts.

3.6 Migrations and Transport

Long-distance migration has evolved in many marine organisms moving through


different habitats and using various modes of locomotion. In most cases migra-
tion has evolved basically as a strategy to maximize fitness in a seasonal envi-
ronment. The basic driving forces for migration are seasonality, spatiotemporal
distributions of resources, habitats, predation and competition. Marine migrants,
moving in an open-ocean environment devoid of visual landmarks, demonstrate
impressive navigational abilities; however the sensory mechanisms used are still
largely unknown. The poor eyesight above water of most of these migrant species
(Ehrenfeld and Koch 1967) probably precludes the use of star patterns and other
celestial cues. Some studies suggest that fronts may act as marks or traces that
3.6  Migrations and Transport 31

can guide nektonic organism in their far ranging migrations across the open ocean.
Westward movement of loggerhead sea turtles (Caretta caretta) across the central
North Pacific occurs along fronts, moving north and south to stay within a spe-
cific frontal zone. Horizontal gradients in temperature, current, chlorophyll, and
possibly prey abundance levels around the fronts may provide cues that logger-
heads would use to maintain their association with fronts (Polovina et al. 2000).
However, other species such as the leatherback turtle (Dermochelys coriacea)
possess a truly remarkable compass sense, allowing them to follow precise tracks
even in the presence of strong currents; in these cases the orientation of their
tracks is independent of fronts (Gaspar et al. 2006).
Many of the large pelagics (e.g. swordfish, tunas, whales) also seem to use
fronts as pathways. For example, swordfish could navigate in a coordinates sys-
tem defined along fronts by isotherms and isolumes. Frontal pathways provide
energy savings in migration, enhanced foraging sites, and with the ability of adults
to place their young in particular locations, an important factor in reproductive
success (Olson 2002). Geographic predictability appears as a needed condition
for front functioning as migratory routes. It is hard to differentiate if fronts pro-
vide navigational clues or feeding opportunities; or both. In any case, perception
of frontal features such as patch contrast or abruptness may be fundamental. The
disciplines of animal behavior and landscape ecology thus become tightly inter-
woven in interpreting boundary function and the response of moving organisms
(Cadenasso et al. 2003b).
At smaller spatial scales, it has been shown that under certain circumstances,
mobile fronts may offer a transport mechanism for some components of the
plankton. Though most of planktonic organisms perform daily vertical migrations
in the water column, they are not able to accomplish migratory movements in
the most energetic horizontal flow fields. In Eastern Boundary Current ecosys-
tems, seasonal wind-driven upwelling brings nutrient-rich water to the surface
along the coast. Fronts develop between the cold waters near the coast and the
warmer offshore waters. As the wind forcing relaxes following coastal upwelling
events, the upwelling fronts move onshore. The low-density surface water moves
shoreward over the upwelled water, forming a convergence zone at the front.
This shoreward-moving front concentrates and transports larvae. This may be an
important mechanism promoting the shoreward migration of larval invertebrates
and fish. The relaxation of winds can bring upwelling fronts to shore periodically,
a process that has been linked to intertidal invertebrate recruitment. In this way
front probability is an important predictor of recruitment of multiple taxa across
the California Current Large Marine Ecosystem. Moreover it appears that owing
to variations in bottom topography and/or coast direction bearing, fronts moving
towards the shore do not impinge everywhere along a coast. This alongshore dif-
ference in the contact of upwelling fronts might cause the observed alongshore
differences in recruitment of intertidal invertebrates (Roughgarden et al. 1991;
Shanks et al. 2000; Woodson et al. 2012). These mechanisms may also account
for larvae to settle jointly, increasing population cohesiveness (see the Larvae
retention section).
32 3  Biology of Fronts

Plume fronts (usually having small-scale and being highly mobile), also may
act concentrating and transporting planktonic larvae. Buoyant plumes may propa-
gate onshore during flood tide transporting high concentrations of planktonic lar-
vae at frontal boundaries. Some estuarine fronts could act as a “larval conduit”
by funneling larvae collected at the front to settlement locations. Since fronts
constrain crossfrontal flow they serve to deflect incident flow, resulting in strong
along-frontal flows which transport larvae collected at the front. These larvae
are likely to settle where the front intersects the shore. If the front is anchored
by a topographic feature then this intersection will be a single stationary point,
and probably a site at which the adult population is concentrated (Largier 1993;
Eggleston et al. 1998).
Chapter 4
Management and Conservation
of Marine Life

Abstract Some species of commercial value, mainly pelagic fishes or squids,


display fidelity to localized high-productivity fronts. Aggregations of organ-
isms in predictable places facilitate fishing operations; moreover fishermen may
easily detect certain kinds of fronts (e.g. thermal fronts) by employing satellite
information, improving fisheries’ efficiency. Abundance of some benthic valuable
resources (e.g. scallops) increase in fronts as well, forming dense and profitable
beds. Species having little or no commercial value also concentrate at fronts; con-
sequently, interactions between fisheries and vulnerable or endangered species are
amplified at fronts: high bycatch rates of large pelagic sharks, sea turtles, marine
mammals, and seabirds characterize some fronts. Fronts also exhibit a potential
to concentrate several types of pollutants (e.g. plastics; oil; heavy metals) at their
surface convergence and in sediments, thus endangering species that make use of
the fronts. Because of potential dangers for marine life, fronts may be considered
as valuable candidates for the implementation of protected areas. The ocean’s
storage of carbon and ability to regulate atmospheric carbon dioxide is crucially
dependent on primary production. Although high phytoplankton standing stocks
do not necessarily imply CO2 sequestration, it has been shown that at least some
fronts do present this type of biogeochemical response. It is expected that cli-
mate change will affect several of the physical forcing processes that generate and
maintain fronts. Variations in the intensity of such forcing will affect key ecologi-
cal processes associated with fronts. Because fronts depend on different forcing,
it is expected that the impact and speed of climate change will vary among fron-
tal types and geographical regions. Fronts play a significant role in the ecology of
seas, and their forcing and properties are likely to change in response to climate
change. Thus, it is suggested that fronts would be ideal sites for early monitoring
and assessment of the dynamics of global state variables.

Keywords Pelagic fisheries · Benthic fisheries · Endangered species bycatch · 


Pollutant concentration  ·  Frontal forcing climate change  · CO2 sequestration

© The Author(s) 2015 33


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_4
34 4  Management and Conservation of Marine Life

4.1 Fisheries

Some species of commercial value display fidelity to localized high-productivity frontal


regions. Once abundant prey patches are located, predators may remain in these discrete
regions for extended periods of time which ultimately leads to their spatial aggrega-
tion favoring fishing operations. Moreover, high densities of non-schooling species (e.g.
swordfish) may occur only where loose aggregations are formed for spawning or feed-
ing, which frequently take place at fronts (Podestá et al. 1993). There is abundant liter-
ature supporting the association of longline fisheries directed to large pelagic predators
(e.g. tunas; swordfish; marlin) and fronts (Etnoyer et al. 2004; Lan et al. 2012; Andrade
2003). It is common practice in several pelagic fisheries to use satellite derived SST charts
to locate thermal fronts. This technology allows fishermen to choose the most profitable
areas to fish. Relationship between fronts and demersal fish fisheries are not so clear. In
the case of Icelandic cod it appears to be a well-established fact among fishermen that
regions of high temperature gradients (i.e. thermal fronts) tend to be indicators of fish
schools. It has been suggested that the food for cod, such as capelin, may aggregate in
thermal fronts; therefore it might be expected that the cod would tend to be found at such
fronts (Brynjarsdóttir and Stefánsson 2004). Several short-finned squids (Illex argentinus
from Patagonia; Todarodes pacificus from Japan; I. illecebrosus from the Northwestern
Atlantic) have been shown to be associated with frontal regions occurring between dif-
ferent water masses, and fishing fleets concentrate efforts on feeding or reproduc-
tive aggregations occurring at these fronts (Gong et al. 1993; Brodziak and Hendrickson
1998; Kiyofuji and Saitoh 2004; Powell et al. 2005; Waluda et al. 2008) (Fig. 4.1). Some

Fig. 4.1  Light generated
by jigging vessels fishing
Argentine short-fin squid
(Illex argentinus) concentrate
along the shelf-break front
in Patagonia on April 23,
2014 (Image generated
by Subprograma de
Sensoramiento Remoto—
INIDEP based on data
from NOAA http://www.
class.noaa.gov)
4.1  Fisheries 35

fisheries targeting benthic resources such as scallops also show clear relations to fronts.
Stocks of Patagonian scallop (Zygochlamys patagonica) are widely distributed over the
western South Atlantic shelf but exploitation is carried out on large, discontinuous, recur-
rently located concentrations that match front locations (Bogazzi et al. 2005). In the west-
ern North Atlantic, the highest concentration of many permanent beds of the Sea Scallops
(Placopecten magellanicus) appears to correspond to areas where physical oceanographic
features such as fronts and gyres may keep larval stages in the vicinity of the spawning
population (Hart and Chute 2004).

4.2 Conservation Issues

High biological productivity at fronts concentrates several kinds of species,


improving their foraging opportunities; consequently, interactions between fish-
eries and vulnerable or endangered species are amplified at fronts. Large marine
vertebrates, such as sea turtles, marine mammals and seabirds, have little or no
commercial value, but become hit, entangled or hooked accidentally by fish-
ing gears that are intended for valuable target species (Lewison et al. 2004). Blue
sharks are commonly caught as bycatch in longlines targeting swordfish and
tuna species in the northeast Atlantic (Queiroz et al. 2012). In the North Atlantic,
leatherback turtle (Dermochelys coriacea) forage along productive fronts where
their main prey, gelatinous plankton, is concentrated. This is also where pelagic,
longline fishing boats aggregate and take a number of sea turtles as bycatch
(Ferraroli et al. 2004). Loggerhead sea turtles (Caretta caretta) travel across the
Pacific between their nesting beaches in Japan and Australia and their foraging
habitat in the eastern Pacific. Juvenile sea turtles travel westward, against pre-
vailing currents, along two convergent thermal fronts. This behavior appears to
explain the highest incidental catch rates of loggerheads in the Hawaiian longline
fishery (Polovina et al. 2000). Around the Azores islands in the Atlantic Ocean,
Caretta caretta juveniles are incidentally caught in drifting longlines that target
swordfish, Xiphias gladius. High catch rates coincide with the presence of thermal
fronts and eddies, which are related to major currents and bathymetric characteris-
tics (Ferreira et al. 2011).
Marine birds exploit preys concentrated at ocean fronts, where fisheries dis-
cards seem to provide a good food source. Around the Brazil-Malvinas Confluence
(a major frontal region in the Southwestern Atlantic), the high concentration
of bird’s preys and the aggregation of fisheries resources (swordfish, tunas and
pelagic sharks) lead to the overlap of industrial fisheries with albatross and pet-
rels. This region has one of the highest levels of incidental capture of albatross and
petrels across global longline fisheries (Jiménez et al. 2011). Along the Patagonian
shelf break front, albatrosses (e.g. wandering albatrosses Diomedea exulans) for-
aging areas overlap with longline fisheries directed to the Patagonian toothfish
(Dissostichus eleginoides) which could lead to increased incidental mortality of
these vulnerable bird species (Xavier et al. 2004).
36 4  Management and Conservation of Marine Life

Fig. 4.2  Fronts are able to


concentrate floating debris. a
Garbage patches in the Gulf
of Maine (Rozalia Project for
a Clean Ocean http://rozali
aproject.blogspot.com.ar/).
b Debris float in the Pacific
Ocean off the coast of Japan
after the Tsunami that struck
the nation on March 11, 2011
(U.S. Navy photo by Steve
White)

Other conservation issue to be taken into account regarding fronts is their


potential to concentrate several types of pollutants at their convergent surface
(Fig.  4.2), thus endangering species that use these areas (Belkin et al. 2009;
Lohmann and Belkin 2014). One of the most ubiquitous and long-lasting changes
to oceans is the accumulation and fragmentation of plastics. Many plastics are
buoyant and remain so until they become waterlogged or accumulates too much
epibiota to float. Mega- and macro-plastics have accumulated in the highest densi-
ties in the Northern Hemisphere, adjacent to urban centers, in enclosed seas and
at fronts (Barnes et al. 2009). In the North Pacific, a large increase in debris den-
sity occurs at the Subtropical Convergence Zone, north of the Hawaiian Islands;
increasing risk for pelagic animals that are preferentially foraging in the same
frontal zone (Pichel et al. 2007). In the German Bight, the distribution of float-
ing objects is driven by winds and surface currents. During calm weather, when
relatively stable fronts form in the bight, floating objects accumulate near these
interfaces (Thiel et al. 2011). The Agulhas Current appears to be a major source
of plastic pollution to the seas off the southwestern Cape Province, South Africa,
where frontal convergences produce the highest concentrations (Ryan 1988).
Fronts can also accumulate non-plastic pollutants (Lohmann and Belkin 2014).
Near the Galician coasts (northwest of Spain), a marked surface front is formed
by the meeting of a poleward current and lower salinity coastal waters. This front
4.2  Conservation Issues 37

concentrated oil patches from the Prestige wreck, impeding their entrance into the
Rías Baixas (Álvarez-Salgado et al. 2006). In the Seto Inland Sea (Japan), tidal
and thermohaline fronts showed elevated concentrations of persistent organochlo-
rines in surface waters, and also in organisms and sediments (Tanabe et al. 1991).
In fact fronts are able to concentrate not only floating pollutants near the surface
but also other pollutants such as heavy metals, which tend to accumulate in the
sediments. For example, the highest concentrations of Cu, Zn, Pb, Cd and Ag in
marine sediments from Gdansk Bay, Poland, occur near the mouth of the Vistula
River. These elements are probably scavenged at the hydrological front by Mn and
Fe oxyhydroxides where mixing of Vistula river water with brackish Baltic Sea
water takes place (Glasby and Szefer 1998).
At smaller spatial scales, in absence of sufficiently strong winds or tides, primary
treated sewage (domestic and industrial) discharged through shoreline and deepwa-
ter outfalls into coastal waters, form visible surface plumes that intrude some kilom-
eters seaward and along the coast from their point of discharge. Such sewage plumes
are lenses of low-salinity waters a few meters deep that overlay high-salinity shelf
waters. Small scale fronts usually develop between plumes and shelf waters where
young fishes may concentrate, as a result of advection at fronts as well as behavioral
responses. Surface sewage plumes therefore affect small-scale (<1 km) patterns of
distribution and density of young fishes and may increase and prolong their exposure
to pollutants that can cause sub-lethal and lethal effects (Gray 1996).
Though not marine fronts properly, internal waves show another example of
pollutant concentration in small scale convergence zones. Tidal currents flowing
off the continental shelf (or across reefs or banks) produce large internal waves
that propagate onshore. Surface current over the waves produce alternating zones
of convergence and divergence, and oil spills (and other flotsams) swept into such
convergences will be trapped there and carried onshore (Shanks 1987).
As a result of the potential for increased survival and growth of planktonic
larvae, together with reduced dispersal, fronts experience increased ecological
resilience due to increased recruitment across a range of taxa from ecosystem
engineers (kelps, corals, barnacles, mussels, scallops) to top predators. In addition,
this predictability makes regions of high front probability particularly amenable
to marine conservation and spatial planning efforts (Etnoyer et al. 2004; Woodson
et al. 2012; Scales et al. 2014). Moreover, because of their role in the amplifica-
tion of interactions between fisheries and endangered species; and their ability to
concentrate pollutants, marine fronts may be seen as highly valuable ecosystems
for wildlife conservation actions such as the implementation of high seas marine
protected areas (Queiroz et al. 2012).

4.3 Climate Change

In marine ecosystems, rising atmospheric CO2 and climate change are associated
with concurrent shifts in temperature, circulation, stratification, strength and direc-
tion of prevailing wind, precipitation, river run-off and groundwater contribution,
38 4  Management and Conservation of Marine Life

nutrient input, oxygen content, and ocean acidification (Rijnsdorp et al. 2009;
Doney et al. 2012). Several of these forcing mechanisms are responsible for the
formation and permanence of fronts, and define main frontal properties, with wide
ranging biological effects. Ocean temperature will follow increases in air tem-
perature, although to a lesser extent, owing to the high heat capacity of seawa-
ter. Shallow areas will exhibit larger temperature increases than deeper waters.
Stratification (resulting from the interplay between temperature and wind mixing)
will also be an important factor in all regions, owing to the effect of stratification
on the vertical fluxes of nutrients and organic matter and consequently on bottom-
up processes. Changes in wind speed and direction not only influence mixing
and water circulation in the open ocean, but also affect the strength of upwellings
within shelf and coastal regions (Rijnsdorp et al. 2009). Alterations in precipita-
tion patterns and subsequent delivery of freshwater, nutrients, and sediment will
affect estuarine productivity. Changes in freshwater flux will affect stratification of
coastal waters impacting also vertical nutrient flux (Scavia et al. 2002). Variations
in the intensity of all these forcing mechanisms will affect key ecological pro-
cesses of fronts like biological production, retention of plankton, concentration
and aggregation of inert materials, bentho-pelagic coupling, etc.
One of the most conspicuous signs of climate change have been recent changes
in the seasonal timing (phenology) of life history events (Thackeray et al. 2010).
Species-specific variation in phenological responses to climate can disrupt the syn-
chrony of ecological interactions and potentially affect community persistence.
The majority of spring and summer events have advanced, and across environ-
ments advances in timing were slowest for secondary consumers, thus increasing
the potential risk of temporal mismatch in key trophic interactions; consequently
future climate warming may exacerbate trophic mismatching, further disrupt-
ing the functioning, persistence and resilience of many ecosystems (Thackeray
et al. 2010). In seasonal fronts (e.g. tidal fronts), or permanent fronts exhibiting
seasonal signals (e.g. estuarine fronts), variations in the seasonality (not only
intensity) of forcing could reduce synchrony between frontal patterns and key
biological processes (feeding; reproduction) aggravating disruption of the ecosys-
tems’ functioning.
Because fronts depend on different forcing processes it is expected that the con-
sequences and speed of climate change vary among frontal types; moreover the
effects of climate change are expected to differ in both magnitude and direction
among geographic areas (Rijnsdorp et al. 2009) so features of the same type of
front could be enhanced in one region and weakened in another.
A matter linking climate change and trophic webs at fronts is CO2 seques-
tration. The ocean’s storage of carbon and ability to regulate atmospheric car-
bon dioxide is crucially dependent on primary production; that is the creation of
organic matter from inorganic nutrients and carbon through photosynthesis. The
photosynthesis predominantly utilizes CO2 dissolved in seawater and so provides
a sink for atmospheric CO2 when organic carbon is transferred to deeper water.
The transport of limiting nutrients to the sunlit surface ocean (the euphotic zone)
plays a central role in controlling primary production. Moreover, vertical fluxes
4.3  Climate Change 39

may transport photosynthetically produced organic matter to the seafloor, where


it is rapidly consumed and incorporated into benthic biomass or trapped in sedi-
ments (Berkenbusch et al. 2011). The convergence of waters at ocean fronts may
result in relatively intense downwelling compared to typical rates in other regions.
These intense vertical velocities may be significant in determining the exchange
rate of heat, carbon dioxide, and other gases between the atmosphere and the deep
ocean (Ferrari 2011).
The hydrography of the Southern Ocean is characterized by a series of zonal
fronts which generally circle the Antarctic continent. Models and observational
studies suggest that the Southern Ocean is an important site for sequestering
atmospheric CO2. Lowest CO2 values concomitant with maxima of chlorophyll
were detected near fronts in the Southern Ocean (Robertson and Watson 1995;
Daly et al. 2001). In the Argentine Sea, the near-shore waters are a source of CO2
to the atmosphere while the midshelf region is a CO2 sink. The transition between
source and sink regions closely matches the location of tidal fronts, suggesting
that phytoplankton blooms near the stratified side of the fronts draw the ocean’s
CO2 to very low levels thus promoting further uptake. At the shelf break front the
CO2 flux to the ocean is the largest (Bianchi et al. 2005, 2009; Schloss et al. 2007).
CO2 sequestration is the result of complex interactions of biological; biogeo-
chemical and oceanographic processes; therefore high phytoplankton standing
stocks do not necessarily imply high CO2 sequestration; however, it has been
shown that at least some fronts are an important element of the ocean-atmosphere
coupled climate system (Ferrari 2011).
Chapter 5
Comparisons of Fronts with Other
Boundaries at Sea

Abstract There are boundaries in the ocean which are not fronts, such as the
pycnocline, the interface of water with the air, the sediments or the ice. These
boundaries pose distinctive biotic and abiotic conditions and are the preferred
living space for certain groups of organisms. Due to the gravitational stratifica-
tion, vertical scales in the sea are highly compressed relative to horizontal scales,
so pycnoclines, and the interfaces between water and sediments; ice; and the
atmosphere are all nearly horizontal edges having much larger horizontal scales
than typical fronts. Most of these interfaces are layers of greatly reduced flows,
weakening plankton dispersion. Life tends to congregate at boundaries and such
non-frontal interfaces are places of locally elevated biological activity. However,
non-frontal interfaces lack mechanisms that persistently bring nutrients to pro-
mote phytoplankton production. Fronts concentrate more biological productivity
in narrower places, and their dynamics appear as more complex, characterized by
intense three dimensional flows. Such complexity could explain the wider range
of ecological properties of fronts compared to other interfaces.

Keywords Thermocline · Halocline · Marine interfaces · Abiotic conditions · 


Plankton concentration  ·  Benthic boundary layer

5.1 The Pycnocline Interface

In several regions of the oceans the water column is stratified, that is, there is a low
density layer overlaying a denser layer. The density change between both layers
is abrupt and is referred to as a pycnocline (Fig. 5.1); most fronts are associated
or connected with pycnoclines (e.g. tidal fronts; estuarine fronts), but pycnoclines
exist independently of fronts. Pycnoclines may be driven by temperature (ther-
moclines) or by salinity (haloclines); or by both. Haloclines are more common
in coastal waters due to the continental runoff, while thermoclines are present
in coastal and offshore environments and are basically driven by the balance of
solar heating and wind mixing. In addition to temperature and salinity pycnoclines

© The Author(s) 2015 41


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_5
42 5  Comparisons of Fronts with Other Boundaries at Sea

Fig. 5.1  Fronts and other boundaries at sea. Greenish areas show regions with high abundance
of organisms. f front. Arrows are currents

also tend to mark the transition of other biologically relevant properties, such as
light intensity and nutrient concentration. In most cases the upper layer is more
illuminated, warmer, fresher and poorer in nutrients. Moreover, pycnoclines fre-
quently separate layers moving at different speeds and in different directions.
Consequently, the ecological changes across a few meters to tens of meters thick
pycnoclines are greater than those across any known front which intersects the sea
surface (Longhurst 1998).
The pycnocline presents special conditions, and its ecological characteristics
differ from that of layers above and below (Longhurst 1998). Pycnoclines are fre-
quently characterized by high nutrient concentrations and low light intensities, and
are inhabited by a diverse “shade flora” encompassing diatoms, dynoflagellates,
and other groups (Longhurst 1998). Internal waves1 traveling along the pycnocline
are observed in some regions. Internal waves are likely to increase turbulent trans-
port of nutrients across the pycnocline and to cause vertical oscillations of phyto-
plankton, thereby increasing the average light intensity they are exposed to (Mann
and Lazier 2006). Zooplankton herbivores are often concentrated as dense thin
layers foraging in these strata of enhanced phytoplankton biomass, though the ver-
tical distribution of zooplankton is complicated by the vertical migrations of some
species across the pycnocline. Such migrant species utilize to their advantage the
contrasting ecological conditions of both levels (Longhurst 1998). Zooplankton-
feeder fishes (e.g. sardines and anchovies) may also be attracted to zooplankton
thin layers because of better feeding opportunities (McManus and Woodson 2012).
There are frequently steep gradients in flow velocity and even reversals in flow
direction associated with pycnoclines; consequently pycnoclines are frequently
layers of no motion or greatly reduced flow (McManus and Woodson 2012).

1 In some cases tidal currents may cause oscillations of the pycnocline and waves that propagate
along the interface.
5.1  The Pycnocline Interface 43

Zooplankton foraging behaviors can then lead to the aggregation into regions
with reduced flows and this may result in a decrease in its horizontal dispersal
(McManus and Woodson 2012).
High gradients of water density and viscosity could act as a physical barrier to
certain organisms attempting to migrate across and could result in the accumula-
tion of organisms within and below pycnoclines (Lougee et al. 2002), but there is
abundant evidence from observations on phyto- and zooplankton to support that
the pycnocline should be considered not only as a way station where sinking par-
ticles may aggregate, but also as a preferred habitat of a characteristic group of
planktonic organisms (Longhurst 1998).

5.2 The Sea Water-Sediment Interface

This interface is a preferred habitat for microorganisms (Fig. 5.1). Respiration in


the water-sediment interface is dominated by bacteria and they play a major role
in the decomposition of organic matter that reaches the sea bed. The supply of
organic matter to the bottom originates from primary production in the upper
layer. Most of the primary production that sinks is remineralized during its descent
and only a small proportion of the surface production arrives to the sea bed; nev-
ertheless, this supply of organic matter is the major determinant of abundance and
growth of deep-sea benthic biota (Turley 2000).
Because the near bottom layer represents a boundary between two oceanic bio-
topes (pelagial and benthal), the animal populations living there belong to diverse
ecological groups. This environment is occupied by endemic species, derived
from downward extensions of pelagic planktonic populations, which are often
seasonal in nature, as well as infaunal species emerging into the water column,
often on diel cycles. The presence of organisms in this layer is determined by two
general factors: organism behavior and boundary layer hydrodynamics (Dauvin
and Vallet 2006).
On the sea floor, a boundary layer of reduced flow extends for tens of centim-
eters above the bottom. In very quiet waters nutritious organic matter accumulates
on the bottom, where it is most easily utilized by animals that ingest sediment
directly. At the other extreme of the bottom boundary layer energy spectrum,
strong currents erode the sediment, leaving little organic matter and making sus-
pension feeding difficult because there is so much inorganic matter in suspen-
sion. For the intermediate stages, a gradient in abundance of suspension-feeding
organisms exists. As the suspension-feeding organisms deplete the concentration
of food particles near the sediment-water interface, moderate tidal currents gen-
erate enough turbulence in the bottom boundary layer to increase the supply of
food to the suspension-feeders by turbulent diffusion. Both current velocity and
bottom roughness influence vertical diffusion, which have the effects of increasing
the transport of food material trough this boundary layer (Legendre et al. 1986;
Mann and Lazier 2006). The weakness or lack of near-bottom currents in lakes has
44 5  Comparisons of Fronts with Other Boundaries at Sea

been proposed as the cause for the poor development of benthos in those habitats
(Nixon 1988), thus highlighting the importance of benthic currents in the ocean.

5.3 The Sea Surface-Atmosphere Interface

Physical, chemical and biological conditions differ greatly between the uppermost
5 cm of the ocean and the water below, so this thin upper layer constitutes a dis-
tinctive biotope (Zaitsev 1997). This layer is occupied by organisms which are
exposed to wind drift because they are fixed to the surface by their own buoyancy
or because they are more or less permanently attached to floating particles; and
also by organisms which stay close to the surface in a more temporary and vari-
able manner (Fig. 5.1) (Hempel and Weikert 1972). In a broad sense the surface
dwelling biota is referred to as neuston (neuston may be divided in different cat-
egories, Zaitsev 1997).
Temperature, wave action and solar radiation seem to be the most important
­abiotic factors for neuston organisms (Hempel and Weikert 1972; Zaitsev 1997);
though in some circumstances heavy rain could adversely affect them by diluting the
surface waters (Holdway and Maddock 1983). Wave action could have a destructive
effect on some neuston organisms; damage of fish eggs and a decrease in length of
diatom chains has been reported (Hempel and Weikert 1972).
During daytime, solar radiation (particularly the high intensities of its ultra-
violet and infrared portions) make the uppermost few centimeters a biotope
quite different from the rest of the water column, where radiation is effectively
decreased by absorption and dispersion. The majority of plankton organisms avoid
the surface during daytime, but representatives of many groups ascend to the sur-
face at night. Under lower radiation situations (e.g. during dim light, at night, in
turbid waters, or at high latitudes) the biotope loses its major characteristic fea-
ture. Under such conditions the neuston community becomes less distinct or even
almost identical to the plankton community of the adjacent strata (Hempel and
Weikert 1972; Holdway and Maddock 1983).
Some inorganic and organic material both from the atmosphere and from the
sea may be concentrated in the air-sea interface. In shallow areas, even benthic
particles will be transported up to the surface by turbulence, and may be fixed
there by surface tension. Coastal areas are also particularly rich in insects, plant
pollen, and seeds transported from land by air drift. Colonies of bacteria may grow
attached to these particles, and also to the bubbles of sea foam and to the surface
(Hempel and Weikert 1972).
In clear, tropical and subtropical waters, the first few centimeters under the sur-
face skin will be particularly poor in food. Low nutrient levels and intense radiation
limit phytoplankton growth. Therefore, despite some concentration of food particles
at the very surface, the general picture of the near surface zone is that of a poor
feeding ground. Certain groups of organisms however, have successfully chosen this
biotope as their habitat: larvae of certain fishes; copepods; euphausids; pteropods;
5.3  The Sea Surface-Atmosphere Interface 45

coelenterates; tunicates; the by-the-wind-sailor (Velella velella) and the Portuguese


man-of-war (Physalia physalis), which have submerged bodies and aerial flota-
tion devices; and insect genus Halobates living on top of the surface film (Beebe
1926; Hempel and Weikert 1972; Holdway and Maddock 1983; Zaitsev 1997). In
upwelling and coastal areas however, the growth of neuston is promoted by the
proliferation of phytoplankton supported by nutrient-rich waters and concentration
of animals in the uppermost layer may exceed the lower layer by several orders of
magnitude (Hempel and Weikert 1972).
During hours of high irradiation, the total number of neuston organisms is
relatively small, the competition for food is rather low, and this fact might com-
pensate somewhat for the paucity of total food supply. Therefore, the few grazers
and predators which are well adapted to the adverse abiotic conditions of the sur-
face stratum could find sufficient food as long as they are not selective; and at the
same time, a refuge from predation during daytime. At night, however, the surface
biotope becomes frequented by potential predators (Hempel and Weikert 1972;
Holdway and Maddock 1983). The neuston connect the sea surface and water col-
umn as planktonic larvae develop and migrate downward, and adult animals visit
the surface to feed and reproduce. Neustonic animals of moderate size (0.2–30 mm)
may be consumed by predators from the aquatic (fish; squids; turtles) and the aerial
(birds) environments. Some neustonic organisms may also consume aerobionts fall-
ing on the sea surface, such as insects (Zaitsev 1997).
The sea-surface is particularly subject to anthropogenic contaminants (oil;
chemicals; metals) and also eutrophication, forming an additional hazard to the
neustonic organisms (Hempel and Weikert 1972; Zaitsev 1997). In some regions
moreover, large quantities of floating tar, plastic, and other debris provide a habitat
for certain kind of neustonic organisms such as isopods, hydroids, and egg masses
(Holdway and Maddock 1983).

5.4 The Ice-Water Interface

Microalgae (mainly benthic diatoms) that develop at the ice-water inter-


face (Fig. 5.1) are responsible for an important part of the productivity of polar
seas, where chlorophyll concentrations in the water column are generally low.
Microalgal growth starts during the initial period of sea ice formation in autumn;
then growth at the ice bottom slows down gradually and stops completely when
very low winter light intensities are reached, though ice algae are markedly shade
adapted (Demers et al. 1986; Flores et al. 2012). The growth of ice algae during
spring and autumn extends the short phytoplankton growing season in the water
column. Unlike phytoplankton, ice algae usually grow under relatively stable
conditions of light, temperature and salinity. Low frequency fluctuations of the
growth of ice algae have been associated with nutrient pulses driven by tides and
atmospheric events; the biological dynamics of the ice-water interface is therefore
coupled to the hydrodynamics of the underlying waters. Physical processes that
46 5  Comparisons of Fronts with Other Boundaries at Sea

determine the rate of upward nutrient transport are found to vary both in space and
time. The boundary layer characteristics depend on the smoothness of the local
and surrounding ice, the regularity of the current regime and other factors related
to the gravitational stability of the surrounding water (Demers et al. 1986).
Melting glacial ice in the ocean may lead to stratification associated with the
freshwater input and is also a source of nutrients, particularly Fe, which is dis-
solved and may fertilize the adjacent ocean. This process may be significant in the
Southern Ocean, which is characterized by low iron concentrations (Statham et al.
2008).
Herbivore zooplankton can feed on ice algae during springtime when water col-
umn productivity is low. These under-ice organisms are believed to represent a link
in the transfer of energy from ice algal production to amphipods to sea-birds and
mammals. Amphipods were observed to swim very close to the underside of the
ice and attach themselves for periods of time, as well as sometimes to enter cracks
and holes in the ice. Copepods and krill larvae are abundant under the ice, which
could be preyed upon by amphipods (Demers et al. 1986; Krapp et al. 2008). Ice
algae accessible from the underside of ice floes constitute an important resource
for larval as well as postlarval Antarctic krill. Observed declines of krill popula-
tions in some sectors of the Southern Ocean are presumably linked to decreasing
recruitment success caused by loss of sea ice habitat. The pronounced presence
of Antarctic krill under the ice highlights its potential as an energy transmitter
between the production of ice algae and the pelagic food web (Flores et al. 2012).

5.5 Fronts Contrasted with the Other Interfaces

Although biological properties of the non-frontal boundaries here accounted for


are not homogeneous (they really are different biotopes), they have in common
their large extensions, presenting much higher spatial coverage than fronts, which
are narrow regions. Moreover, non-frontal boundaries are quasi-horizontal instead
of vertically inclined interfaces, that is, they do not develop along the direction
of the gravity force and light that have a central ecological significance at sea
(Margalef 1997). There is a lack of nutrient pumping mechanisms at most non-
frontal boundaries; however some pycnoclines may oscillate due to the passage of
internal waves, producing vertical displacements of nutrients and phytoplankton.
Most of non-frontal boundaries act as a “substrate” that passively may aggregate
or concentrate nutrients and organisms due to physical properties (density, viscos-
ity) or species behavior. However, such concentrations may be locally important.
Due to the lesser spatial scale and the action of mechanisms generating upward
nutrient fluxes, fronts are characterized by much higher concentration of biologi-
cal productivity than other interfaces. As mentioned above, the mechanical energy
of the ocean becomes available for biological production at fronts. This auxiliary
energy is not directly used by plants, but it is efficient in providing a more or less
constant supply of nutrients for phytoplankton.
Chapter 6
Comparisons of Fronts with Terrestrial
Boundaries and the “Ecotone” Concept

Abstract  Marine fronts possess most of the main characteristics of ecotones, as


defined for land ecosystems, so fronts can be considered as true ecotones. In addi-
tion to land ecotones features, marine fronts are characterized by high biological
productivity, a property rarely (if ever) reported for terrestrial ecotones. Moreover,
concentration and retention of small planktonic forms in marine fronts plays a
much more important ecological role than analogous processes on land. Both high
biological production and retention/concentration are responsible for the notewor-
thy concentration of life at marine fronts. On the other hand, terrestrial landscapes
may be modified by human actions, usually increasing contrasts across boundaries
(habitat fragmentation); this is another significant difference with marine fronts,
which cannot be created by human actions, though they may present significant
long-term variability in response to climate change. Fronts and ecotones can rep-
resent unique habitats optimal to some species and inhospitable to others; they are
places of tension where evolution forces may be at work. The age and history of
fronts and ecotones may determine their functional properties, their ecological
impacts increasing with their persistence.

Keywords Ecotones ·  Environmental gradients  ·  Biological production  · Habitat


fragmentation  ·  Atmospheric fronts  ·  Marine fronts

Definitions of fronts refer us to the “ecotone” concept. As is true for landscape


ecology in general, the study of boundaries has largely been directed to terrestrial
systems (Cadenasso et al. 2003b; Sournia 1994), thus it remains to be explored if
marine fronts can be considered as genuine ecotones.
Ecotones can be viewed as zones where spatial or temporal rates of change in
ecological structure or function are rapid relative to rates across the landscape as a
whole (Hansen et al. 1992). A broad definition provided by di Castri and Hansen
(1992) states that an ecotone is “a zone of transition between adjacent ecological
systems, having a set of characteristics uniquely defined by space and time scales,
and by the strength of the interactions between adjacent ecological systems”.

© The Author(s) 2015 47


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_6
48 6  Comparisons of Fronts with Terrestrial Boundaries …

Landscapes resemble mosaics composed by two kinds of structures: patches


and boundaries. Ecotones or boundaries are important structural features in
the landscape; they represent areas of transition, contact or separation between
the contrasting elements of a mosaic. Patches in the landscapes are connected
by fluxes of organisms, materials, energy, and information; and although eco-
tones are likely to occupy a relatively restricted part of the total volume, they are
expected to control the interactions of mosaics elements by modulating the con-
necting flows, thus becoming important control points (Cadenasso et al. 2003a, b).
Landscapes may be seen as dynamic systems of patches that are coalescing or dis-
integrating primarily through the expansion, fluctuation or erosion of boundaries
through time. Although some boundaries may be relatively stable in ecological
time, in other cases, interactions within biotic transitions will results in landscapes
characterized by fluctuations in the number, type and size of patches, along with
directional shifts from one patch type to another in response to local and regional
drivers (Peters et al. 2006). In terms of system dynamics, ecotones are locations
where the rates of ecological transfers (e.g. energy flow, nutrient exchange) change
abruptly in relation to those within the homogeneous units on either side. The
view of the oceans as a complex of different water masses looks analogous to the
perception of landscapes as mosaics, and the quoted general properties of e­ cotones
closely resemble that of the marine fronts. Following Gosz (1992), emergent
­properties of the ecotones are more likely to occur where there are:
(1) Contrasting life history strategies or life forms on each side of the transition
(this is a common pattern in opposite sides of several marine frontal types,
for example the dominance of diatoms or dynoflagellates in each side of tidal
fronts; or that of neritic or oceanic planktonic forms across shelf-break fronts);
(2) Different constraints operating on organisms of each biome (contrasting phys-
icochemical properties in several fronts comprise an acute set of constrains,
for example light versus osmotic constrains at each side in estuarine fronts;
temperature versus nutrients for plants in shelf-break fronts; light versus
­nutrients at each side in tidal fronts);
(3) Different scale-related features in the two biomes (because of differences in
temperature, planktonic organisms may have different generation time at each
side of a front);
(4) Different heterogeneity features (for example vertical homogeneity versus
stratification of the water column present in each side of tidal fronts or shelf-
break fronts; different types of bottom substrates at each side of estuarine or
tidal fronts).
Though some boundary habitats in land (e.g. saltmarshes) have been reported as
hotspots of biological production (Canepuccia et al. 2011), biological productiv-
ity does not seem to be a general characteristics of terrestrial ecotones; in contrast,
marine fronts concentrate a disproportionate portion of the biological production
at sea and this is a main reason for their ecological importance.
Ecotones may serve either as barriers or corridors between gene pools, but this
does not necessarily imply that ecotones show higher diversity of species as
6  Comparisons of Fronts with Terrestrial Boundaries … 49

compared with neighboring environments.1 For instance, highly fluctuating


­boundaries (in space or time) will be relatively poor in species diversity (di Castri
and Hansen 1992). Several measures of biodiversity have been proposed
(Whittaker 1972); alpha diversity refers to the diversity within a particular area or
ecosystem; gamma diversity is a measure of the overall diversity for the different
ecosystems within a region; and beta diversity represents the turnover of species
between habitats or localities. Gamma and alpha diversity would not necessarily
exhibit a pattern relative to ecotones, even though the species identities would
clearly change across an ecotone; in these cases beta diversity should increase
toward the ecotone (Neilson et al. 1992). It has been shown that copepods living in
an area of high oceanographic complexity present peaks in beta diversity associ-
ated with fronts (Berasategui et al. 2006). Although the effect of fronts on larger,
free-swimming animals is more difficult to quantify, it was shown that the pres-
ence of fronts influences fish diversity, with dissimilar effects for different frontal
types (Alemany et al. 2009). Fronts could also be distributional barriers for some
species but at the same time to concentrate others increasing diversity.
Thermoregulation, optimal foraging, and behavioral bioenergetics could each lead
to fish aggregations at fronts or could cause a front to appear as a barrier to the
distribution of a fish species (Brandt 1993). It has also been proposed that the pat-
terns of biodiversity across ecotones may become clearer if attributes of ecotones
structure (e.g. width; contrast) are normalized by a relevant spatial scale (body
size, territory size, or species range), or time scale (age until first breeding,
­longevity, dispersal rate) (Hansen et al. 1992).
It has been predicted that conditions in the ecotones may be a simple average of
conditions in the patches on either side, or alternatively that they may reflect inter-
actions that occur along the boundary (Strayer et al. 2003). The physical matrix of
most fronts matches the first proposition (e.g. horizontal pattern of temperature in
a tidal front, or that of salinity in an estuarine front); however ecological proper-
ties of fronts such as biological production, beta diversity or species abundances
are clearly better reflected by the second prediction. Ecotones are places of tension
where stresses of a genetic character, important in evolution, are at work (Margalef
1997). This could be the case of some frontal types, for example estuarine fronts:
because life originated in the sea, the colonization of the freshwaters must have
occurred, partially at least, through fronts separating the freshwaters and the estua-
rine waters, so the new capabilities and forms have evolved in response to those
fronts.
Because the effects of a boundary may be cumulative, the age and history of
an ecotone may determine its functional properties and the local ecological con-
ditions around the boundary (Strayer et al. 2003). The biological importance of
fronts depends on a matching of physical and biological time scales. Noting the
longer times scales at higher trophic levels and the successive nature of increasing

1  Inthe case of fronts the issue of the relationship between diversity and productivity also needs
to be considered. See the Diversity section.
50 6  Comparisons of Fronts with Terrestrial Boundaries …

trophic levels, one can expect that fronts which persist for longer times will be
characterized by the presence of higher trophic levels. Thus, the presence of large
predators like big fishes are unlikely at fronts with short time scales, more typical
of small estuaries for example (Brandt 1993; Largier 1993). On the other hand,
the phytoplankton landscape may be organized into patches of around 10–100 km,
often dominated by a particular phytoplankton group, separated by physical fronts
induced by horizontal stirring. These physical fronts effectively delimit ephemeral
ecological niches by encircling water masses of similar history and whose life-
times are comparable to the timescale of the phytoplankton biological response
(a few weeks) (d’Ovidio et al. 2010).
Specific locations in a landscape can serve as a boundary for one research ques-
tion and as a patch for a different question (Cadenasso et al. 2003a). Marine fronts
share this dual character with terrestrial ecotones; they may also be considered
to function as distinct ecosystems, in so far as the physicochemical environment
and the biological characteristics found in a frontal region may differ markedly
from those of adjacent waters (Sournia 1994; Polovina et al. 2001). Ecotones can
represent unique habitats optimal to some species and inhospitable to others (di
Castri and Hansen 1992); for example, a biome transition zone is hypothesized
to have properties different from adjacent biomes and may amplify or attenu-
ate some system processes such as productivity, resource dynamics and avail-
ability (Gosz 1992). Likewise, the physics of fronts provides unique opportunities
for various types of organisms; while at the same time, they can lead to an acute
set of physiological challenges. Some fauna would use fronts as prime foraging
grounds making use of the fact that some prey are at a disadvantage in the front
because of thermal, haline or nutritional stresses (Olson 2002). Human modifica-
tions of terrestrial landscapes overlying natural environmental heterogeneity (habi-
tat fragmentation) is resulting in an increase in the number and types of ecological
patterns and their intervening boundaries (Peters et al. 2006), moreover some ter-
restrial boundaries may be experimentally modified in order to test scientific
hypotheses. This is a substantial difference with marine fronts, which cannot be
created by humans; although fronts’ properties can be altered; for example climate
change may play a role on modifying the location and perhaps the existence of
some fronts.
Much of our current theoretical understanding of marine fronts arises from
meteorological sciences (Olson 2002). Because of their fluid nature, the atmos-
phere and the oceans are governed by the same fundamental physical laws, and
consequently both systems present similar dynamics and analogous structures.
Flying insects and other components of the “aerial plankton” can be concentrated
by wind convergences (Russell 1999; Chapman et al. 2011). Large-scale atmos-
pheric fronts are responsible for assisting the transport of insects into coastal
regions and landscape-induced sea/lake breeze circulations provide a mechanism
to explain the accumulations of a wide variety of insects on shorelines of large
water bodies (Isard et al. 2001). Flying insects concentrated by atmospheric fronts
and convergence of air masses settle and may attract insectivorous birds (Russell
1999) in analogous way to zooplankton aggregations at marine fronts attract
6  Comparisons of Fronts with Terrestrial Boundaries … 51

fishes and other predators. Many atmospheric fronts are produced at terrestrial
ecotones, such as shorelines of large water bodies or the land/sea boundary (Isard
et al. 2001); however they are hardly seen as a vertical extension of such ecotones.
Moreover, in contrast with marine fronts, the location of few atmospheric fronts
is locked to the topography of land and they therefore displace thousands of kilo-
meters. Although the resemblance of the physical dynamics of atmospheric and
marine fronts is remarkable, the ecological effects of marine fronts are of greater
consequence. This is so because primary producers in the sea are planktonic
­(phytoplankton) and their abundances are strongly influenced by flows at fronts:
vertical flows that bring nutrients into the illuminated upper layer and convergent
flows that concentrate and retain phytoplankton at frontal interfaces. Moreover,
primary consumers in the sea (a key element in the energy transference to upper
trophic levels) are mostly planktonic (e.g. copepods) and are also severely influ-
enced by frontal dynamics. Finally, most of the fishes and invertebrates possess
planktonic larval stages. In the atmosphere, there are no species equivalents to
marine phytoplankton; consequently the atmospheric circulation is not relevant for
nutrient distribution and atmospheric fronts cannot promote primary production.
In addition, the flying insects concentrated at atmospheric fronts are adult stages
and play no role in the trophic webs comparable to that of marine zooplankton.
Temporal scales are also different, while atmospheric fronts last for hours or days
(Steele 1991) most marine fronts are permanent or seasonal; as mentioned above,
the ecological impact of ecotones increase with their persistence.
Chapter 7
Final Remarks

Abstract  Simple facts like the differences in density and light absorption between
air and water make terrestrial and marine ecosystems qualitatively different. The
knowledge on how terrestrial systems work and their comparative analysis with
the sea, has greatly improved our understanding of the functioning of the oceans.
Auxiliary energy is vital for primary production in the sea, and marine fronts are
places where this kind of energy becomes available for life processes, making
fronts fundamentally different from terrestrial ecosystems and more complex than
terrestrial ecotones. Of course, fronts are not the only places in the oceans that
provide auxiliary energy for life, but they are relatively small areas that concen-
trate a disproportionate quantity of such energy. Fluid processes that make availa-
ble auxiliary energy are also responsible for the rest of the ecological properties of
fronts, including facilities such as retention and landmarks; all these features make
fronts sites of utmost ecological importance, they are the oases of the oceans.

Keywords Marine migrations cues · “Meeting point” hypothesis · Plankton


retention  ·  Fish reproduction grounds  ·  Marine ecosystems productivity  · Plant
nutrients upwelling  ·  Climate change  ·  Mechanical energy redistribution

Fronts are structural and functional components of the seascape. They may affect
biological production, community structure, biogeography patterns, life his-
tory processes, and provide signals and pathways for pelagic species. In relation
to man, they are important for fisheries and conservation, and affected by pollu-
tion and global change. Marine fronts lead to conditions that allow organisms to
accomplish their vital challenges. They provide two essential facilities: landmarks
in a traceless realm and mechanical energy for plankton retention and biological
production.

© The Author(s) 2015 53


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4_7
54 7  Final Remarks

7.1 Landmarks and Beacons: Orienting and Meeting

The pelagic environment is characterized by its extensiveness, homogeneity, lack


of refuge and seasonality and patchiness of food availability. Migration is a com-
mon strategy for coping with seasonality in the production cycle. The question of
how migrating animals can find their way while covering immense distances, have
long fascinated and puzzled mankind. Migrating in the ocean poses formidable
challenges to marine animals because the open ocean appears as a most uniform
environment, lacking visual cues that are known to be fundamental in guiding
migrations on land. Olson (2002) proposed that fronts may provide clues for large-
scale migrations. Fronts are narrow regions which represent anomalies in the flow
field and display high contrast in physical as well as biological variables, so their
detection and tracking could be feasible for some migratory species, assisting
migrants to follow a course during their journeys. Because of the high prey abun-
dance at fronts, individuals can feed while travelling making it difficult to disen-
tangle foraging and migration processes.
Group formation is a widespread phenomenon throughout the animal kingdom.
Different reasons for grouping include decreasing the predation risk, promoting
optimal foraging, increasing reproductive success and facilitating migration and
learning. The homogeneous and expansive nature of the marine pelagic domain
favors dispersion. Hence, for gregarious fish species some processes must enhance
the gathering of individuals (Soria et al. 2009). Several fishes (notably tuna) living
in a rather uniform universe are attracted by any physical “anomaly” in their envi-
ronment (e.g. floating objects, bottom discontinuity, high gradients). The “meet-
ing point hypothesis” Freón and Dagorn (2000) considers that schooling pelagic
species can make use of floating objects to increase the encounter rate between
isolated individuals or small schools and other schools. For an isolated fish, or a
small group of fish, the floating object might be easier to detect than a large school
(using one or several senses: vision, audition, olfaction). In addition, from the
point of view of the energy balance, waiting for conspecifics at a meeting point
is less costly than swimming to find them. The ability of fronts to concentrate
residue floating at the sea surface (e.g. kelp, branches or trunks of tree, vegetal
debris), and any industrial residue (e.g. boards, housing, parts of destroyed fishing
gears) could make them useful meeting points for pelagic fishes.

7.2 Mechanical Energy for Retention

The ocean is a highly dispersive environment. Dispersion is a combination of


advection and eddy diffusion. Advection is the mean transport of a collection of
particles, while diffusion reflects the differences in transport of individual parti-
cles. In words of Largier (2003), imagine “a cloud of dye in the water: it moves
7.2  Mechanical Energy for Retention 55

Fig. 7.1  The ocean
is a highly dispersive
environment. Dispersion is
a combination of advection
and eddy diffusion. Dots
represent floating particles
(e.g. plankton) at the
beginning (t1) and at the
finish (t2) of an observation
period

along and spreads out as it does so—advection is the mean movement and dif-
fusion is the spreading out” (Fig. 7.1). Dispersion poses two problems for drift-
ing and mobile organisms: they are constantly being transported by advection
well beyond their “normal” ranges, into localities where they may persist but not
breed; and at the same time scattered in different directions by diffusion, persis-
tently increasing the distance to the nearest mate, and diminishing the encoun-
ter probability among conspecifics; this has vital consequences for those species
with sexual reproduction. Sinclair (1988) clearly stated the critical role of spa-
tial processes for the establishment and persistence of marine populations. The
emergence of sex during life evolution brought with the constraint of physical
relationship between individuals; in other words, the advantage of sexual mode
of reproduction was gained at the expense of freedom of individual in geographic
space. Consequently, the retention of the individuals of a population in relatively
fixed geographic space, increasing the probability of sexual encounter, becomes
a crucial challenge. Population persistence for planktonic species requires appro-
priate encounter rates during sexual reproduction, and retention processes within
fronts are important in this respect. Nektonic species with a larval stage show a
more complicated situation, in which adult homing to natal retention area is
involved, and with frontal dynamics playing a significant role for the establish-
ment of reproductive grounds. Mechanical and thermal energy are responsible
for water dynamics. At fronts, such dynamics offer opportunities for planktonic
organisms (including larvae of fish and invertebrates) to be retained. There are
frequently steep gradients in flow velocity and even reversals in flow direction
associated with fronts. Tuning between circulation and vertical migratory behav-
ior permits retention of planktonic organisms in favorable ecological locations
and at the same time, increases the cohesiveness of individuals of the popula-
tion. This is why fronts play a role in the establishment of populations’ patterns of
marine organisms.
56 7  Final Remarks

Fig. 7.2  Energy sources for biological production at sea. Solar radiation is the source of primary
energy for photosynthesis, the rest are processes that supply auxiliary energy for marine ecosystems

7.3 Mechanical Energy for Biological Production

In the entire biosphere, the primary source of energy for ecosystems is the pho-
tosynthetically active radiation of the sun; but in the oceans, such energy needs
to be complemented. Legendre et al. (1986) and Margalef (1997) conceptualized
the matter: besides this primary energy, the productivity of marine ecosystems
depends on the input of mechanical energy derived from de degradation of solar
energy (e.g. winds, freshwater runoff, air-ocean heat exchanges), or of energy of
gravitational nature (tides). This auxiliary (or exosomatic) energy (Fig. 7.2) is not
directly used by living organisms, but it is efficient in increasing the storage of
solar energy by the phytoplankton which is then transferred along the food web.
Auxiliary energy on proper scales makes possible the replenishment of the lim-
iting plant nutrients, and thus biological production. Marine fronts are important
because at these locations the auxiliary energy becomes available for life pro-
cesses. The biological production in the sea is unevenly distributed; even with its
immense productivity the oceans have relatively low production across much of
their domain. The importance of the auxiliary energy is evident when we realize
that the spatio-temporal patterns in marine biological production are much more
related to the spatio-temporal distributions of auxiliary energy than to those of pri-
mary energy. This is of paramount importance also because global change is redis-
tributing auxiliary energy in both space and time (Jumars et al. 2009); and this is
7.3  Mechanical Energy for Biological Production 57

why the effects of global change may be different in terrestrial systems and in the
oceans. On land, nutrient transport is internalized and under control of the vascular
plants, whereas in the oceans it depends on the auxiliary energy. Consequently,
ocean fronts seem to be ideal sites for early monitoring and assessment of global
change effects.
Acknowledgments  We are grateful to Mike Sinclair and Don Olson whose criticism and useful
comments on an earlier draft have considerably improved this manuscript. We are grateful also
to Mara Braverman for the artwork. This research was supported by grants from UNMdP EXA
555/12, and PIP 112-201101-00892 to E.M.A.; and from the Inter-American Institute for Global
Change Research (IAI) grants CRN2076 and CRN 3070 sponsored by the US National Science
Foundation Grant GEO-0452325 and GEO-1128040. This is INIDEP Contribution no. 1895.
Literature Cited

Abrams PA (1995) Monotonic or unimodal diversity productivity gradients: what does competition
theory predicts? Ecology 76:2019–2027
Acha EM, Mianzan H, Guerrero R, Favero M, Bava J (2004) Marine fronts at the continental
shelves of austral South America. Physical and ecological processes. J Mar Syst 44:83–105
Alemany D, Acha EM, Iribarne O (2009) The relationship between marine fronts and fish diversity
in the SW Atlantic shelf large marine ecosystem. J Biogeogr 36:2111–2124
Álvarez-Salgado XA, Herrera JL, Gago J, Otero P, Soriano JA, Pola CG, García-Soto C (2006)
Influence of the oceanographic conditions during spring 2003 on the transport of the Prestige
tanker fuel oil to the Galician coast. Mar Pollut Bull 53:239–249
Andrade HA (2003) The relationship between the skipjack tuna (Katsuwonus pelamis) fishery and
seasonal temperature variability in the south-western Atlantic. Fish Oceanogr 12(1):10–18
Angel MV (1986) Vertical distribution: study and implications. In: Pierrot-Bults AC, van der Spoel
S, Zahuranec BJ, Johnson RK (eds) Pelagic biogeography. Proceedings of an international
conference, The Netherlands, 29 May-5 June 1985 (UNESCO, Technical Papers in Marine
Science N° 49, France, pp 3–8)
Arístegui J, Sangra P, Hernández-León S, Cantón M, Hernández-Guerra A, Kerling JL (1994)
Island-induced eddies in the Canary Islands. Deep Sea Res Part I Oceanogr Res Pap
41(10):1509–1525
Ashjian CJ (1993) Trends in copepod species abundances across and along a Gulf Stream meander:
evidence for entrainment and detrainment of fluid parcels from the Gulf Stream. Deep Sea Res
I 40(3):461–482
Bakun A (1996) Patterns in the ocean: ocean processes and marine population dynamics. Sea
Grant, La Jolla
Bakun A (2006a) Fronts and eddies as key structures in the habitat of marine fish larvae: opportu-
nity, adaptive response and competitive advantage. Sci Mar 70S2:105–122
Bakun A (2006b) Wasp-waist populations and marine ecosystem dynamics: Navigating the preda-
tor pit topographies. Prog Oceanogr 68:271–288
Barnes DKA, Galgani F, Thompson RC, Barlaz M (2009) Accumulation and fragmentation of
plastic debris in global environments. Philos Trans Roy Soc London B 364:1985–1998
Barth JA, Bogucki D, Pierce SD, Kosro PM (1998) Secondary circulation associated with a
shelfbreak front. Geophys Res Lett 25(15):2761–2764
Bates M (1960) The forest and the sea: a look at the economy of nature and the ecology of man.
Random House, New York
Beebe W (1926) The Arcturus adventure. G.P. Putnam’s Sons, New York
Belkin IM, Cornillon PC, Sherman K (2009) Fronts in large marine ecosystems. Prog Oceanogr
81:223–236

© The Author(s) 2015 59


E.M. Acha et al., Ecological Processes at Marine Fronts,
SpringerBriefs in Environmental Science, DOI 10.1007/978-3-319-15479-4
60 Literature Cited

Berasategui AD, Menu Marque S, Gómez-Erache M, Ramírez FC, Mianzan HW, Acha EM
(2006) Copepods assemblages in a highly complex hydrographic region. Estuar Coast Shelf
Sci 66:483–492
Berkenbusch K, Probert PK, Nodder SD (2011) Comparative biomass of sediment benthos across a
depth transect, Chatham Rise, Southwest Pacific Ocean. Mar Ecol Prog Ser 425:79–90
Bernatchez L, Martin S (1996) Mitochondrial DNA diversity in anadromous rainbow smelt,
Osmerus mordax Mitchill: a genetic assessment of the member-vagrant hypothesis. Can J
Fish Aquat Sci 53(2):424–433
Bianchi A, Bianucci L, Piola A, Ruiz Pino D, Schloss I, Poisson A, Balestrini CF (2005)
Vertical stratification and air-sea CO2 fluxes in the Patagonian shelf. J Geophys Res
110(C07003):1–10
Bianchi AA, Ruiz Pino D, Isbert Perlender HG, Osiroff AP, Segura V, Lutz V, Clara ML,
Balestrini CF, Piola AR (2009) Annual balance and seasonal variability of sea-air CO2 fluxes
in the Patagonia Sea: their relationship with fronts and chlorophyll distribution. J Geophys
Res 114:C03018
Bogazzi E, Baldoni A, Rivas AL, Martos P, Reta R, Orensanz JM, Lasta ML, Dell Arciprete
OP, Werner F (2005) Spatial correspondence between areas of concentration of Patagonian
scallop (Zygochlamys patagonica) and frontal systems in the southwestern Atlantic. Fish
Oceanogr 14(5):359–376
Boltovskoy D, Correa N, Boltovskoy A (2005) Diversity and endemism in cold waters of the
South Atlantic: contrasting patterns in the plankton and the benthos. Sci Mar 69(2):17–26
Bost CA, Cotté C, Cherel Y, Charrassin JB, Guinet C, Ainley DG, Weimerskirch H (2009) The
importance of oceanographic fronts to marine birds and mammals of the southern oceans.
J Mar Syst 78:363–376
Bowman MJ (1978) Introduction and historical perspectives. In: Bowman MJ, Esaias WE (eds)
Oceanic fronts in coastal processes. Springer, New York, pp 2–5
Brandini F, Boltovskoy D, Piola AR, Kocmur S, Rottgers R, Abreu P, Mendes Lopes R (2000)
Multiannual trends in fronts and distribution of nutrients and chlorophyll in the southwestern
Atlantic. Deep Sea Res I 47:1015–1033
Brandt SB (1983) Pelagic squid associations with a warmcore eddy of the East Australian current.
Aust J Mar Freshw Res 34:573–585
Brandt SB (1993) The effect of thermal fronts on fish growth: a bioenergetics evaluation of food
and temperature. Estuaries 16(1):142–159
Breder CM, Rosen DE (1966) Modes of reproduction in fishes. The Natural History Press, New
York
Brodziak J, Hendrickson L (1998) An analysis of environmental effects on survey catches of
squids Loligo pealei and Illex illecebrosus in the northwest Atlantic. Fish Bull 97:9–24
Brooks SC (1934) Oceanic currents and the migration of pelagic birds. Condor 36(5):185–190
Brynjarsdóttir J, Stefánsson G (2004) Analysis of cod catch data from Icelandic groundfish surveys
using generalized linear models. Fish Res 70:195–208
Cadenasso ML, Pickett STA, Weathers KC, Bell SS, Benning TL, Carreiro MM, Dawson TE
(2003) An interdisciplinary and synthetic approach to ecological boundaries. BioScience
53(8):717–722
Cadenasso ML, Pickett STA, Weathers KC, Jones CG (2003) A framework for a theory of
­ecological boundaries. BioScience 53(8):750–758
Campagna C, Piola A, Marín MR, Lewis M, Fernández T (2006) Southern elephant seal trajectories,
fronts and eddies in the Brazil/Malvinas confluence. Deep Sea Res I 53:1907–1924
Canepuccia AD, Montemayor D, Pascual J, Farina JL, Iribarne OO (2011) A stem-boring moth
drives detritus production in SW Atlantic marshes. Mar Ecol Prog Ser 442:1–9
Castro JJ, Santiago JA, Santana-Ortega AT (2002) A general theory on fish aggregation to floating
objects: an alternative to the meeting point hypothesis. Rev Fish Biol Fish 11:255–277
Cury P, Shannon L, Shin Y-J (2001) The functioning of marine ecosystems. In: Reykjavik conference
on responsible fisheries in the marine ecosystem, Reykjavik, Iceland, 1–4 Oct 2001
Cushman Murphy R (1944) Wet lands and dry seas. Nat Hist 53(8):350–356
Literature Cited 61

Chapman JW, Drake VA, Reynolds DR (2011) Recent insights from radar studies of insect flight.
Annu Rev Entomol 56:337–356
D’Asaro E, Lee C, Rainville L, Harcourt R, Thomas L (2011) Enhanced turbulence and energy
dissipation at ocean fronts. Science 332:318–322
Daly KL, Smith WO Jr, Johnson GC, Di Tullio GR, Jones DR, Mordy CW, Feely RA, Hansell
DA, Zhang J-Z (2001) Hydrography, nutrients, and carbon pools in the Pacific sector of the
Southern Ocean: implications for carbon flux. J Geophys Res 106:7107–7124
Darwin C (1845) Journal of researches into the natural history and geology of the countries vis-
ited during the voyage of H M S. Beagle round the world, under the command of Captain Fitz
Roy. R.N. John Murray, London, UK, 519 pp
Dauvin J-C, Vallet C (2006) The near-bottom layer as an ecological boundary in marine ecosys-
tems: diversity, taxonomic composition and community definitions. Hydrobiologia 555:49–58
Demers S, Legendre L, Therriault JC, Ingram RG (1986) Biological production at the ice-water
ergocline. In: Nihoul JCJ (ed) Marine interfaces ecohydrodynamics. Elsevier oceanography
series, vol 42. Elsevier, New York, 31–54
Dewicke A, Rottiers V, Mees J, Vincx M (2002) Evidence for an enriched hyperbenthic fauna in
the Frisian front (North Sea). J Sea Res 47:121–139
Díaz Briz L, Martorelli SR, Genzano GN, Mianzan H (2012) Parasitism (Trematoda, Digenea)
in medusae from the southwestern Atlantic Ocean: medusa hosts, parasite prevalences, and
ecological implications. Hydrobiologia 690(1):215–226
di Castri F, Hansen AJ (1992) The environment and development crises as determinants of land-
scape dynamics. In: Hansen AJ, di Castri F (eds) Landscape boundaries. Consequences for biotic
diversity and ecological flows. Springer, New York, pp 3–18
Doney SC, Ruckelshaus M, Duffy JM, Barry JP, Chan F, English CA, Galindo HM, Grebmeier
JM, Hollowed AB, Knowlton N, Polovina J, Rabalais NN, Sydeman WJ, Talley LD (2012)
Climate change impacts on marine ecosystems. Ann Rev Mar Sci 4:11–37
Dong C, Mavor T, Nencioli F, Jiang S, Uchiyama Y, McWilliams JC, Dickey T, Ondrusek M,
Zhang H, Clark DK (2009) An oceanic cyclonic eddy on the lee side of Lanai Island, Hawaii.
J Geophys Res Oceans 114(C10):C10008
Dong C, McWilliams JC, Shchepetkin AF (2007) Island wakes in deep water. J Phys Oceanogr
37(4):962–981
d’Ovidio F, De Monte S, Alvain S, Dandonneau Y, Lévy M (2010) Fluid dynamical niches of
phytoplankton types. Proc Nat Acad Sci USA 107(43):18366–18370
Ducklow HW, Steinberg DK, Buesseler KO (2001) Upper ocean carbon export and the biological
pump. Oceanography 14:50–58
Dunson WA, Ehlert GW (1971) Effects of temperature, salinity, and surface water flow on distri-
bution of the sea snake Pelamis. Limnol Oceanogr 16(6):845–853
Dutkiewicz S, Follows MJ, Bragg JG (2009) Modeling the coupling of ocean ecology and bio-
geochemistry. Global Biogeochem Cycles 23:GB4017
Eggleston DB, Armstrong DA, Elis WE, Patton WS (1998) Estuarine fronts as conduits for lar-
val transport: hydrodynamics and spatial distribution of Dungeness crab postlarvae. Mar Ecol
Prog Ser 164:73–82
Ehrenfeld DW, Koch AL (1967) Visual accommodation in the green turtle. Science 155:827–828
Eppley RW, Peterson BJ (1979) Particulate organic matter flux and planktonic new production in
the deep ocean. Nature 282:677–680
Etnoyer P, Canny D, Mate B, Morgan L (2004) Persistent pelagic habitats in the Baja California
to Bering Sea (B2B) ecoregion. Oceanography 17(1):90–101
Fedorov KN (1986) The physical nature and structure of oceanic fronts, vol 19. Lectures notes on
coastal and estuarine studies. Springer, New York
Fenchel T (1988) Marine plankton food chains. Ann Rev Ecol Syst 19:19–38
Ferrari R (2011) A frontal challenge for climate models. Science 332:316–317
Ferraroli S, Georges J-Y, Gaspar P, Le Maho Y (2004) Where leatherback turtles meet fisheries.
Nature 429:521–522
62 Literature Cited

Ferreira RL, Martins HR, Bolten AB, Santos MA, Erzini K (2011) Influence of environmental
and fishery parameters on loggerhead sea turtle by-catch in the longline fishery in the Azores
archipelago and implications for conservation. J Mar Biol Assoc UK 91(8):1697–1705
Flores H, van Franeker JA, Siegel V, Haraldsson M, Strass V, Meesters EH, Bathmann U, Wolff
WJ (2012) The association of Antarctic krill Euphausia superba with the under-ice habitat.
PLoS ONE 7(2):1–11
Franklin B (1786) A letter from Dr. Benjamin Franklin to Mr. Alphonsus le Roy containing sundry
marine observations. Trans Am Philos Soc 2:294–329
Franks PJS (1992) Sink or swim: accumulation of biomass at fronts. Mar Ecol Prog Ser 82:1–12
Freón P, Dagorn L (2000) Review of fish associative behaviour: toward a generalisation of the
meeting point hypothesis. Rev Fish Biol Fish 10:183–207
Frontier S (1986) Studying fronts as contact ecosystems. In: Nihoul JCJ (ed) Marine interfaces
ecohydrodynamics, vol 42. Elsevier oceanography series. Elsevier, New York, pp 55–66
Gaard E (1996) Life cycle, abundance and transport of Calanus finmarchicus in faroese waters.
Ophelia 44:59–70
Gaard E, Gislason A, Falkenhaug T, Soiland H, Musaeva E, Vereshchaka A, Vinogradov G
(2008) Horizontal and vertical copepod distribution and abundance on the Mid-Atlantic
Ridge in June 2004. Deep Sea Res Part II 55:59–71
Gaertner JC, Bertrand JA, Gil de Sola L, Durbec JP, Ferrandis E, Souplet A (2005) Large spa-
tial scale variation of demersal fish assemblage structure on the continental shelf of the NW
Mediterranean Sea. Mar Ecol Prog Ser 297:245–257
Gaines SD, Lester SE, Eckert G, Kinlan BP, Sagarin R, Gaylord B (2009) Dispersal and geographic
ranges in the sea. In: Witman JD, Roy K (eds) Marine macroecology. The University of Chicago
Press, Chicago, pp 227–249
Garvine RW (1975) The distribution of salinity and temperature in the Connecticut River estuary.
J Geophys Res 80:1176–1183
Gaspar P, Georges J-Y, Fossette S, Lenoble A, Ferraroli S, Le Maho Y (2006) Marine animal
behaviour: neglecting ocean currents can lead us up the wrong track. Proc Roy Soc London B
273:2697–2702
Gaston KJ (2000) Global patterns in biodiversity. Nature 405:220–227
Gawarkiewicz G, Chapman DC (1992) The role of stratification in the formation and mainte-
nance of shelf-break fronts. J Phys Oceanogr 22:753–772
Gaylord B, Gaines SD (2000) Temperature or transport? Range limits in marine species mediated
solely by flow. Am Nat 155:769–789
Glasby GP, Szefer UP (1998) Marine pollution in Gdansk Bay, Puck Bay and the Vistula Lagoon,
Poland: an overview. Sci Total Environ 212:49–57
Gong Y, Kim S, An DH (1993) Abundance of neon flying squid in relation to oceanographic con-
ditions in the North Pacific. Int North Pac Fish Comm Bull 53:191–204
Gosz JR (1992) Ecological functions in a biome transition zone: translating local responses to
broad-scale dynamics. In: Hansen AJ, di Castri F (eds) Landscape boundaries. Consequences
for biotic diversity and ecological flows. Springer, New York, pp 55–75
Govoni JJ, Grimes CB (1992) The surface accumulation of larval fishes by hydrodynamic con-
vergence within the Mississippi River plume front. Cont Shelf Res 12(11):1265–1329
Graham RT, Witt MJ, Castellanos DW, Remolina F, Maxwell S, Godley BJ, Hawkes LA (2012)
Satellite tracking of Manta rays highlights challenges to their conservation. PLoS ONE
7(5):e36834
Gray CA (1996) Intrusions of surface sewage plumes into continental shelf waters: interactions
with larval and presettlement juvenile fishes. Mar Ecol Prog Ser 31:31–45
Haney JC (1986) Seabird segregation at Gulf Stream frontal eddies. Mar Ecol Prog Ser 28:279–285
Hansen AJ, Risser PG, di Castri F (1992) Epilogue: biodiversity and ecological flows across ecotones.
In: Hansen AJ, di Castri F (eds) Landscape boundaries. Consequences for biotic diversity and
ecological flows. Springer, New York, pp 423–438
Hart DR, Chute AS (2004) Essential Fish Habitat Source Document. NOAA Technical Memorandum,
32 pp
Literature Cited 63

Hempel G, Weikert H (1972) The neuston of the subtropical and boreal North-eastern Atlantic
Ocean. Rev Mar Biol 13:70–88
Holdway P, Maddock L (1983) A comparative survey of neuston: geographical and temporal dis-
tribution patterns. Mar Biol 76:263–270
Hosia A, Stemmann L, Youngbluth M (2008) Distribution of net-collected planktonic cnidarians
along the northern Mid-Atlantic Ridge and their associations with the main water masses.
Deep Sea Res Part II 55:106–118
Houde ED (2009) Recruitment variability. In: Jakobsen T, Fogarty MJ, Megrey BA, Moksness
E (eds) Fish reproductive biology. Implications for assessment and management. Wiley,
Chichester, pp 91–171
Houghton RW, Visbeck M (1998) Upwelling and convergence in the Middle Atlantic Bight shelf-
break front. Geophys Res Lett 25(15):2765–2768
Hunt BPV, Hosie GW (2003) The Continuous Plankton Recorder in the Southern Ocean: a com-
parative analysis of zooplankton communities sampled by the CPR and vertical net hauls
along 140 E. J Plankton Res 25(12):1561–1579
Huyer A (1983) Coastal upwelling in the California Current system. Prog Oceanogr 12:259–284
Iles TC, Sinclair M (1982) Atlantic herring: stock discreteness and abundance. Science
215:627–633
Isard SA, Kristovich DAR, Gage SH, Jones CJ, Laird NF (2001) Atmospheric motion systems
that influence the redistribution and accumulation of insects on the beaches of the Great
Lakes in North America. Aerobiologia 17:275–291
Jackson GA, Checkley JDM (2011) Particle size distributions in the upper 100 m water column
and their implications for animal feeding in the plankton. Deep Sea Res I 58:283–297
Jeffrey SW, Hallegraeff GM (1980) Studies of phytoplankton species and photosynthetic pig-
ments in a warmcore eddy of the East Australian Current. I. Summer population. Mar Ecol
Prog Ser 3:285–294
Jiménez S, Domingo A, Abreu M, Brazeiro A (2011) Structure of the seabird assemblage associ-
ated with pelagic longline vessels in the southwestern Atlantic: implications for bycatch. Mar
Ecol Prog Ser 15:241–254
John H-C, Mohrholz V, Lutjeharms JRE (2001) Cross-front hydrography and fish larval distribu-
tion at the Angola-Benguela Frontal Zone. J Mar Syst 28:91–111
Johnson GC, McPhaden MJ, Firing E (2001) Equatorial Pacific ocean horizontal velocity, diver-
gence, and upwelling. J Phys Oceanogr 31(3):839–849
Jumars PA, Trowbridge JH, Boss E, Karp-Boss L (2009) Turbulence-plankton interactions: a new
cartoon. Mar Ecol 30:133–150
Kiorboe T (1993) Turbulence, phytoplankton cell size, and the structure of pelagic food webs.
Adv Mar Biol 29:1–72
Kiorboe T (2008) A mechanistic approach to plankton ecology. Princeton University Press,
Princeton
Kiyofuji H, Saitoh S-I (2004) Use of nighttime visible images to detect Japanese common squid
Todarodes pacificus fishing areas and potential migration routes in the Sea of Japan. Mar Ecol
Prog Ser 276:173–186
Klein P, Lapeyre G (2009) The oceanic vertical pump induced by mesoscale and submesoscale
turbulence. Ann Rev Mar Sci 1:351–375
Krapp RH, Berge J, Flores H, Gulliksen B, Werner I (2008) Sympagic occurrence of Eusirid and
Lysianassoid amphipods under Antarctic pack ice. Deep Sea Res II 55:1015–1023
Lalli CM, Parsons TR (1997) Biological oceanography. An Introduction. Butterworth-
Heinemann, Oxford
Lan K-W, Kawamura H, Lee M-A, Lu H-J, Shimada T, Hosoda K, Sakaida F (2012) Relationship
between albacore (Thunnus alalunga) fishing grounds in the Indian Ocean and the thermal
environment revealed by cloud-free microwave sea surface temperature. Fish Res 113:1–7
Largier JL (1993) Estuarine fronts: how important are they? Estuaries 16(1):1–11
Largier JL (2003) Considerations in estimating larval dispersal distances from oceanographic
data. Ecol Appl 13(1):71–89
64 Literature Cited

Le Févre J (1986) Aspects of the biology of frontal systems. Adv Mar Biol 23:163–299
Le Févre J, Frontier S (1988) Influence of temporal characteristics of physical phenomena on
plankton dynamics, as shown by North-West European marine ecosystems. In: Rothschild BJ
(ed) Toward a theory on biology-physical interactions in the World Ocean. Kluwer Academic
Publishers, Dordrecht, pp 245–272
Legendre L, Demers S, Lefaivre D (1986) Biological production at marine ergoclines. In: Nihoul
JCJ (ed) Marine interfaces ecohydrodynamics, vol 42. Elsevier oceanography series. Elsevier,
New York, pp 1–29
Leichter JJ, Witman JD (2009) Basin-scale oceanographic influences on marine macroecologi-
cal patterns. In: Witman JD, Roy K (eds) Marine macroecology. The University of Chicago
Press, Chicago, pp 205–226
Levin LA (2006) Recent progress in understanding larval dispersal: new directions and digressions.
Integr Comp Biol 46(3):282–297
Lévy M, Ferrari R, Franks PJS, Martin AP, Riviére P (2012) Bringing physics to life at the
submesoscale. Geophys Res Lett 39:L14602
Lévy M, Klein P, Tréguier AM (2001) Impact of submesoscale physics on production and subduc-
tion of phytoplankton in an oligotrophic regime. J Mar Res 59:535–565
Lewison RL, Crowder LB, Read AJ, Freeman SA (2004) Understanding impacts of fisheries
bycatch on marine megafauna. Trends Ecol Evol 19(11):598–604
Lohmann R, Belkin IM (2014) Organic pollutants and ocean fronts across the Atlantic Ocean: a
review. Prog Oceanogr 128:172–184
Longhurst A (1998) Ecological geography of the sea. Academis Press, San Diego
Lougee LA, Bollens SM, Avent SR (2002) The effects of haloclines on the vertical distribution
and migration of zooplankton. J Exp Mar Biol Ecol 278:111–134
Lucifora LO, García VB, Menni RC, Worm B (2012) Spatial patterns in the diversity of sharks,
rays, and chimaeras (Chondrichthyes) in the Southwest Atlantic. Biodivers Conserv 21:407–419
Mahadevan A, D’Asaro E, Lee C, Perry MJ (2012) Eddy-driven stratification initiates North
Atlantic spring phytoplankton blooms. Science 337(6090):54–58
Mann KH, Lazier JRN (2006) Dynamics of marine ecosystems. Biological-physical interactions
in the oceans. Blackwell Science Publications, Cambridge
Margalef R (1978) Life-forms of phytoplankton as survival alternatives in an unstable environment.
Oceanol Acta 1(4):493–509
Margalef R (1997) Our biosphere. Excellence in ecology series. Ecology Institute Luhe, Germany
Matano R, Palma ED (2008) On the upwelling of downwelling currents. J Phys Oceanogr
38:2482–2500
Matano RP, Palma ED (2010) The upstream spreading of bottom-trapped plumes. J Phys
Oceanogr 40(7):1631–1650
Mauna AC, Acha EM, Lasta ML, Iribarne OO (2011) The influence of a large SW Atlantic shelf-
break frontal system on epibenthic community composition, trophic guilds and diversity. J
Sea Res 66(1):39–46
McGillicuddy DJJ, Robinson AR, Siegel DA, Jannasch HW, Johnson R, Dickey TD, McNeil
J, Michaels AF, Knap AH (1988) Influence of mesoscale eddies on new production in the
Sargasso Sea. Nature 394:263–266
McGinty N, Power AM, Johnson MP (2011) Variation among northeast Atlantic regions in the
responses of zooplankton to climate change: Not all areas follow the same path. J Exp Mar
Biol Ecol 400:120–131
McManus MA, Woodson CB (2012) Plankton distribution and ocean dispersal. J Exp Biol
215(6):1008–1016
Moteki M, Koubbi P, Pruvost P, Tavernier E, Hulley P-A (2011) Spatial distribution of pelagic fish
off Adélie and George V Land, East Antarctica in the austral summer 2008. Polar Sci 5:211–224
Munk W, Wunsch C (1998) Abyssal recipes II: energetics of tidal and wind mixing. Deep Sea
Res I 45:1977–2010
Naylor E (2006) Orientation and navigation in coastal and estuarine zooplankton. Mar Freshw
Behav Physiol 39(1):13–24
Literature Cited 65

Neilson RP, King GA, DeVelice RL, Lenihan JM (1992) Regional and local vegetation patterns: the
responses of vegetation diversity to subcontinental air masses. In: Hansen AJ, di Castri F (eds)
Landscape boundaries. Consequences for biotic diversity and ecological flows. Springer, New
York, pp 129–149
Nevitt G (1999) Olfactory foraging in Antarctic seabirds: a species-specific attraction to krill
odours. Mar Ecol Prog Ser 177:235–241
Nixon SW (1988) Physical energy inputs and the comparative ecology of lake and marine eco-
systems. Limnol Oceanogr 33(4 Part 2):1005–1025
Norris RD (2000) Pelagic species diversity, biogeography, and evolution. Paleobiology
26(4):236–258
O’Donnell J (1993) Surface fronts in estuaries: A review. Estuaries 16:12–39
Ohman MD, Powell JR, Picheral M, Jensen DW (2012) Mesozooplankton and particulate matter
responses to a deep-water frontal system in the Southern California current system. J Plankton
Res 34(9):815–827
Olson DB (2002) Biophysical dynamics of ocean fronts. In: Robinson AR, McCarthy JJ,
Rothschild BJ (eds) The sea. Biological-physical interactions in the sea, vol 12. Wiley, New
York, pp 187–218
Olson DB, Backus RH (1985) The concentrating of organisms at fronts: a cold-water fish and a
warm-core Gulf Stream ring. J Mar Res 43:113–137
Olson DB, Hitchcock GL, Mariano AJ, Ashjian CJ, Peng G, Nero RW, Podestá GP (1994) Life
on the edge: marine life and fronts. Oceanography 7(2):52–60
Orsi AH, Whitworth T III, Nowlin WD Jr (1995) On the meridional extent and fronts of the
Antarctic Circumpolar Current. Deep Sea Res Part I 42(5):641–673
Ortner PB, Hulbert EM, Wiebe PH (1979) Phytohydrography, Gulf Stream rings, and herbivore
habitat contrasts. J Exp Mar Biol Ecol 39:101–124
Ortner PB, Wiebe PH, Haury LR, Boyd SH (1978) Variability in zooplankton biomass distribu-
tion in the northern Sargasso Sea: the contribution of Gulf Stream cold-core rings. Fish Bull
76:323–334
Pauly D, Christensen V (2005) Primary production required to sustain global fisheries. Nature
374:255–257
Peters DPC, Gosz JR, Pockman WT, Small EE, Parmenter RR, Collins SL, Muldavin E (2006)
Integrating patch and boundary dynamics to understand and predict biotic transitions at mul-
tiple scales. Landscape Ecol 21:19–33
Piatt JF, Springer AM (2003) Advection, pelagic food webs and the biogeography of seabirds in
Beringia. Mar Ornithol 31:141–154
Pichel WG, Churnside JH, Veenstra TS, Foley DG, Friedman KS, Brainard RE, Nicoll JB, Zheng
Q, Clemente-Colón P (2007) Marine debris collects within the North Pacific Subtropical
Convergence Zone. Mar Pollut Bull 54:1207–1211
Pingree RD, Pugh PR, Holligan PM, Forster GR (1975) Summer phytoplankton blooms and red
tides along tidal fronts in the approaches to the English Channel. Nature 258:672–677
Podestá GP (1990) Migratory pattern of Argentine Hake Merluccius hubbsi and oceanic pro-
cesses in the Southwestern Atlantic Ocean. Fish Bull 88(1):167–177
Podestá GP, Browder JA, Hoey JJ (1993) Exploring the association between swordfish catch and
thermal fronts on the U.S. longline grounds in the western North Atlantic. Cont Shelf Res
13:253–277
Polovina J, Kobayashi DR, Parker DM, Seki MP, Balazs GH (2000) Turtles on the edge: move-
ment of loggerhead turtles (Caretta caretta) along oceanic fronts, spanning longline fishing
grounds in the central North Pacifc, 1997–1998. Fish Oceanogr 9:71–82
Polovina JF, Howell E, Kobayashi DR, Seki MP (2001) The transition zone chlorophyll front,
a dynamic global feature defining migration and forage habitat for marine resources. Prog
Oceanogr 49:469–483
Powell EN, King SE, Bonner AJ (2005) Determinants of temporal trends in size in vessel-based
reporting in the USA Illex illecebrosus fishery. J Appl Ichthyol 21:184–197
66 Literature Cited

Queiroz N, Humphries NE, Noble LR, Santos AM, Sims DW (2012) Spatial dynamics and
expanded vertical niche of blue sharks in oceanographic fronts reveal habitat targets for con-
servation. PLoS ONE 7(2 e32374):1–12
Reynolds CS (2006) Ecology of phytoplankton. Cambridge University Press, Cambridge
Reynolds RW, Smith TM, Liu C, Chelton DB, CK S, Schlax MG (2007) Daily high-resolution-
blended analyses for sea surface temperature. J Clim 20:5473–5496
Rijnsdorp AD, Peck MA, Engelhard GH, Möllmann C, Pinnegar JK (2009) Resolving the effect
of climate change on fish populations. ICES J Mar Sci 66:1570–1583
Robert M, Dagorn L, Lopez j, Moreno G, Deneubourg J-L (2013) Does social behavior influence
the dynamics of aggregations formed by tropical tunas around floating objects? An experi-
mental approach. J Exp Mar Biol Ecol 440:238–243
Robertson JE, Watson AJ (1995) A summer-time sink for atmospheric carbon dioxide in the
Southern Ocean between 88°W and 80°E. Deep Sea Res Part II 42(4–5):1081–1091
Rodriguez J, Tintoré J, Allen JT, Blanco JM, Gomis D, Reul A, Ruiz JVR, Echevarria F, Jiménez-
Gómez F (2001) Mesoscale vertical motion and the size structure of phytoplankton in the
ocean. Nature 410:360–363
Roosenburg WM (1996) Maternal condition and nest choice: an alternative for the maintenance
of environmental sex determination? Am Zool 36:157–168
Rosenzweig ML, Abramsky Z (1993) How are diversity and productivity related? In: Ricklefs
RE, Schluter D (eds) Species diversity in ecological communities: historical and geographical
perspectives. University of Chicago Press, Chicago, pp 52–65
Roughgarden J, Pennington JT, Stoner D, Alexander S, Miller DK (1991) Collisions of upwell-
ing fronts with the intertidal zone: the cause of recruitment pulses in barnacle populations of
central California. Acta Oecologica 12:35–51
Russell RW (1999) Precipitation scrubbing of aerial plankton: inferences from bird behavior.
Oecologia 118:381–387
Russell RW, Harrison NM, Hunt GL Jr (1999) Foraging at a front: hydrography, zooplankton,
and avian planktivory in the northern Bering Sea. Mar Ecol Prog Ser 182:77–93
Ryan PG (1988) The characteristics and distribution of plastic particles at the sea-surface of the
Southwestern Cape Province, South Africa. Mar Environ Res 25:249–273
Sánchez-Velasco L, Lavín MF, Jiménez-Rosenberg SPA, Montes JM, Turk-Boyer PJ (2012)
Larval fish habitats and hydrography in the Biosphere Reserve of the Upper Gulf of
California (June 2008). Cont Shelf Res 33:89–99
Scales KL, Miller PI, Hawkes LA, Ingram SN, Sims DW, Votier SC (2014) On the front line:
frontal zones as priority at-sea conservation areas for mobile marine vertebrates. J Appl Ecol
51(6):1575–1583
Scavia D, Field JC, Boesch DF, Buddemeier RW, Burkett V, Cayan DR, Fogarty M, Harwell MA,
Howarth RW, Mason C, Reed DJ, Royer TC, Sallenger AH, Titus JG (2002) Climate change
impacts on US coastal and marine ecosystems. Estuaries 25(2):149–164
Schloss IR, Ferreyra GA, Ferrario ME, Almandoz GO, Codina R, Bianchi AA, Balestrini CF,
Ochoa HA, Ruiz Pino D, Poisson A (2007) Role of plankton communities in sea–air varia-
tions in pCO2 in the SW Atlantic Ocean. Mar Ecol Prog Ser 332:93–106
Shanks AL (1987) The onshore transport of an oilspill by internal waves. Science
235(4793):1198–1200
Shanks AL, Largier JL, Brink L, Brubaker J, Hooff R (2000) Demonstration of the onshore
transport of larval invertebrates by the shoreward movement of an upwelling front. Limnol
Oceanogr 45(1):230–236
Sheldon RW, Sutcliffe JWH, Paranjape MA (1977) Structure of pelagic food chain and relation-
ship between plankton and fish production. J Fish Res Board Can 34:2344–2353
Simpson JH, Hunter JR (1974) Fronts in the Irish Sea. Nature 250:404–406
Sims DW, Southall EJ, Tarling GA, Metcalfe JD (2005) Habitat-specific normal and reverse diel
vertical migration in the plankton-feeding basking shark. J Anim Ecol 74:755–761
Sinclair M (1988) Marine populations. An essay on population regulation and speciation.
University of Washington Press, Seattle
Literature Cited 67

Smetacek V (1999) Revolution in the ocean. Nature 401:647


Sokolov S, Rintoul SR (2009) Circumpolar structure and distribution of the Antarctic Circumpolar
Current fronts: 2. Variability and relationship to sea surface height. J Geophys Res 114:C11019
Sommer U, Stibor H, Katechakis A, Sommer F, Hansen T (2002) Pelagic food web con-
figurations at different levels of nutrient richness and their implications for the ratio fish
production:primary production. Hydrobiologia 484:11–20
Soria M, Dagorn L, Potin G, Freón P (2009) First field-based experiment supporting the meeting
point hypothesis for schooling in pelagic fish. Anim Behav 78:1441–1446
Sournia A (1994) Pelagic biogeography and fronts. Prog Oceanogr 34:109–120
Spalding MD, Agostini VN, Rice J, Grant SM (2012) Pelagic provinces of the world: a biogeo-
graphic classification of the world’s surface pelagic waters. Ocean Coast Manag 60:19–30
Spalding MD, Fox HE, Allen GR, Davidson N, Ferdaña ZA, Finlayson M, Halpern BS, Jorge
MA, Lombana A, Lourie SA, Martin KD, Mcmanus E, Molnar J, Recchia CA, Robertson
J (2007) Marine ecoregions of the world: a bioregionalization of coastal and shelf areas.
BioScience 57(7):573–583
Statham PJ, Skidmore M, Tranter M (2008) Inputs of glacially derived dissolved and colloi-
dal iron to the coastal ocean and implications for primary productivity. Global Biogeochem
Cycles 22:GB3013
Steele JH (1989) The ocean ‘landscape’. Landscape Ecol 3(3/4):185–192
Steele JH (1991) Can ecological theory cross the land-sea boundary? J Theor Biol 153:425–436
Stemmann L, Youngbluth M, Robert K, Hosia A, Picheral M, Paterson H, Ibañez F, Guidi L,
Lombard F, Gorsky G (2008) Global zoogeography of fragile macrozooplankton in the upper
100–1000 m inferred from the underwater video profiler. ICES J Mar Sci 65:433–442
Strathmann RR (1990) Why life histories evolve differently in the sea. Am Zool 30:197–207
Strayer DL, Power ME, Faga WF, Pickett STA, Belnap J (2003) A classification of ecological
boundaries. BioScience 53(8):723–729
Sverdrup HU (1953) On conditions for the vernal blooming of phytoplankton. J Conseil Int
l’Explor de la Mer 18:287–295
Takeoka H, Kaneda A, Anami H (1997) Tidal fronts induced by horizontal contrast of vertical
mixing efficiency. J Oceanogr 53:563–570
Tanabe S, Nishimura A, Hanaoka S, Yanagi T, Takeoka H, Tatsukawa R (1991) Persistent organo-
chlorines in coastal fronts. Marine Pollution Bulletin 22(7):344–351
Taylor AG, Goericke R, Landry MR, Selph KE, Wick DA, Roadman MJ (2012) Sharp gradients
in phytoplankton community structure across a frontal zone in the California current ecosys-
tem. J Plankton Res 34(9):778–789
Taylor JR, Ferrari R (2011) Ocean fronts trigger high latitude phytoplankton blooms. Geophys
Res Lett 38(L23601):1–5
Tesi T, Langone L, Ravaioli M, Giglio F, Capotondi L (2012) Particulate export and lateral
advection in the Antarctic Polar Front (Southern Pacific Ocean): one-year mooring deploy-
ment. J Mar Syst 105–108:70–81
Thackeray SJ, Sparksw TH, Frederiksenz M, Burthe S, Bacon PJ, Bellk JR, Botham MS,
Breretonw TM, Bright PW, Carvalho L, Clutton-Brock T, Dawson A, Edwards MN, Elliott
JM, Harrington R, Johns D, Jones ID, Jones JT, Leech DI, Roy DB, Scott WA, Smith M,
Smithers RJ, Winfield IJ, Wanless S (2010) Trophic level asynchrony in rates of phenological
change for marine, freshwater and terrestrial environments. Glob Change Biol 16:3304–3313
Thiel M, Hinojosa IA, Joschko T, Gutow L (2011) Spatio-temporal distribution of floating
objects in the German Bight (North Sea). J Sea Res 65:368–379
Tranter DJ, Tafe DJ, Sandland RL (1983) Some zooplankton characteristics of warm-core eddies
shed by the East Australian Current, with particular reference to copepods. Aust J Mar
Freshw Res 34:587–607
Turley C (2000) Bacteria in the cold deep-sea benthic boundary layer and sediment-water inter-
face of the NE Atlantic. FEMS Microbiol Ecol 33:89–99
Uda M (1938) Research on shiome or current rip in the seas and oceans. Geophysl Mag
11:307–372
68 Literature Cited

Vlietstra LS, Coyle KO, Kachel NB, Hunt JL Jr (2005) Tidal front affects the size of prey used
by a top marine predator, the short-tailed shearwater (Puffinus tenuirostris). Fish Oceanogr
14(1):196–211
Waluda CM, Griffiths HJ, Rodhouse PG (2008) Remotely sensed spatial dynamics of the Illex
argentinus fishery, Southwest Atlantic. Fish Res 91:196–202
Warner RR, Cowen RK (2002) Local retention of production in marine populations: evidence,
mechanisms, and consequences. Bull Mar Sci 70(1 Suppl.):245–249
Weinstein MP, Weiss SL, Hodson RG, Gerry LR (1980) Retention of three taxa of postlarval
fishes in an intensively flushed tidal estuary, Cape Fear River, North Carolina. Fish Bull
78:419–436
Whittaker RH (1972) Evolution and measurement of species diversity. Taxon 21(2/3):213–251
Wolanski E, Hamner WM (1988) Topographically controlled fronts in the ocean and their bio-
logical influence. Science 241:177–181
Woodson CB, McManus MA, Tyburczy JA, Barth JA, Washburn L, Caselle JE, Carr MH, Malone
DP, Raimondi PT, Menge BA, Palumbi SR (2012) Coastal fronts set recruitment and connec-
tivity patterns across multiple taxa. Limnol Oceanogr 57(2):582–596
Worm B, Lotze HK, Myers RA (2003) Predator diversity hotspots in the blue ocean. Proc Nat
Acad Sci USA 100(17):9884–9888
Wright DH, Currie DJ, Maurer BA (1993) Energy supply and patterns of species richness on
local and regional scales. In: Ricklefs RE, Schluter D (eds) Species diversity in ecological
communities: historical and geographical perspectives. University of Chicago Press, Chicago,
pp 66–74
Wyrtki K, Kilonsky B (1984) Mean water and current structure during the Hawaii-to-Tahiti shut-
tle experiment. J Phys Oceanogr 14(2):242–254
Xavier JC, Trathan PN, Croxall JP, Wood AG, Podestá G, Rodhouse PG (2004) Foraging ecol-
ogy and interactions with fisheries of wandering albatrosses (Diomedea exulans) breeding at
South Georgia. Fish Oceanogr 13(5):324–344
Zaitsev Y (1997) Neuston of seas and oceans. In: Liss PS, Duce RA (eds) The sea surface and
global change. Cambridge University Press, Cambridge, pp 371–382

You might also like