Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

SCI. MAR., 61 (Supl.

1): 159-176 SCIENTIA MARINA 1997

LECTURES ON PLANKTON AND TURBULENCE, C. MARRASÉ, E. SAIZ and J.M. REDONDO (eds.)

Turbulence and ichthyoplankton: influence on vertical


distributions and encounter rates*
SVEIN SUNDBY
Institute of Marine Research. P.O. Box 1870 Nordnes, N-5024 Bergen, Norway.

SUMMARY: Two different aspects of the effects of turbulent mixing on eggs and fish larvae are considered here. In the
first topic the effects of physical processes on the vertical distribution of the eggs are considered. The physical processes
which determine the vertical distribution of fish eggs, and all other immobile plankton, are the buoyancy of the plankton and
the vertical mixing of the water column. Analytical models are presented to demonstrate the influence of the various terms.
A numerical model is also applied to show the effects of non-stationary solutions. The model results are compared with field
measurements. In the second topic the effects of turbulence on plankton encounter rates are considered. The processes are
illustrated by field data on first-feeding cod larvae feeding on Calanus nauplii. These field data show that wind-induced tur-
bulence strongly enhances the feeding rate of the larvae. For first-feeding cod larvae the feeding rate increases by a factor
of about 7 when the wind speed increases from 2 m s-1 to 10 m s-1. Model simulations show how the influence of turbulence-
enhanced encounter rate decreases as the larvae grow, increase their swimming speed and switch to larger prey of lower con-
centration. The simulations show that the turbulence-induced encounter rate decreases exponentially as cod grow, but sig-
nificantly influences feeding until the stage of 2 months old juveniles.

Key words: Fish eggs, buoyancy, vertical mixing, fish larvae, feeding rate, wind-induced turbulence, encounter rate.

INTRODUCTION understanding the effects of turbulence on plankton-


ic life many aspects of the theoretical framework are
Physical processes influence planktonic life in a still insufficiently explored. Mann and Lazier (1991)
large variety of ways, but only a handful of basic gave an overview on turbulence from an ecological
physical parameters are involved in these processes. perspective.
The three most important parameters are tempera- Traditionally, the detrimental effects of turbu-
ture, light and turbulence. A large literature exists on lence have been given focus in ichtyoplankton
effects of temperature and light. These parameters research. Much attention has been given especial-
are easy to measure and relatively easy to simulate ly to how turbulence by vertical mixing breaks
and control in experimental designs. Turbulence, down patchy high concentrations of food, which
however, is difficult to measure, particularly in field in turn is supposed to decrease feeding and sur-
experiments, and only during recent years accept- vival of fish larvae (e.g. Lasker, 1975). Turbu-
able sensors for turbulence measurements have been lence, however, affects plankton dynamics in dif-
developed. More importantly because of the lack of ferent ways, with both positive effects on plankton
production, as well as detrimental effects. Cury
and Roy (1989) studied pelagic fish recruitment in
*Received December 5, 1995. Accepted April 17, 1996. upwelling areas. They found that there was an

TURBULENCE AND ICHTHYOPLANKTON 159


optimal level of wind velocity with respect to fish VERTICAL DISTRIBUTIONS OF FISH EGGS
recruitment, and suggested that this was caused by
the opposing effects of wind-induced upwelling, Basic equations
which enhances plankton production, and turbu-
lent mixing, which has detrimental effects on lar- Let us assume that the horizontal variation in egg
val feeding. Studies on grazing copepods indicate distribution can be neglected compared to the verti-
that turbulence may have multiple effects on the cal variations. The vertical component of the diffu-
feeding of zooplankton, particularly linked to sion equation is then the basic equation in describ-
behavioral responses to turbulence (Alcaraz et al. ing the vertical distribution of eggs:
1988; Saiz and Alcaraz, 1992). Consequently, dur-
ing recent years we have become increasingly ∂ C(z, t) ∂ [ w(z, t)C(z, t)]
aware of the adverse effects of turbulence. For this – =
∂t ∂z
reason, proxy variables for turbulence should be
used with care. Investigations of biological ∂ ⎡ ∂ C(z,t) ⎤
= K(z,t) + S(z,t) + M(z,t) (1)
responses and effects of turbulence should prefer- ∂ z ⎢⎣ ∂ z ⎥⎦
ably be made by studying the basic biological and
physical processes. where
Before exploring how biological processes are
influenced by turbulence, it is important to realise C(z,t) concentration of eggs in numbers per
that light and turbulent mixing are the most promi- unit volume
nent physical features of the upper layer of the sea, w(z,t) vertical velocity of the eggs
the major site for all forms of planktonic produc- K(z,t) vertical eddy diffusivity coefficient
tion. Obviously, light is the basic energy source S(z,t) spawning (production) of eggs
for synthesizing biomass. However, turbulence in M(z,t) egg mortality
the upper ocean, induced by wind stress, transfers
energy from the atmosphere to the euphotic zone, If we neglect the influence of spawning and egg
and turbulence is the most important means for mortality on the vertical distribution and consider
transferring momentum from the large scale to the stationary conditions, equation (1) reduces to
small scale. Therefore, from a general point of
view, it should not be surprising that turbulent ∂ C(z)
energy also contributes to planktonic biomass pro- –w(z) ⋅ C(z) = K(z) (2)
∂z
duction.
Principally, there are two quite different aspects
of how turbulence influences plankton. Firstly, tur- The equation expresses the balance between the
bulence influences plankton distributions, as for vertical velocity of the eggs (left hand side) and the
example might be described theoretically by Fick- eddy diffusion flux of the eggs (right hand side).
ian diffusion. This is a eulerian framework using Equation (2) can be solved to give the vertical
eddy exchange coefficients to describe the effects distribution of eggs. The solution depends on the
of turbulence. Secondly, turbulence influences the shape of the functions describing vertical velocity,
encounter rate between individual planktonic parti- w(z), and eddy diffusivity, K(z). Let us, therefore,
cles, i. e. predators and prey, which is most conve- first consider the factors which determine the values
niently described in a lagrangeian framework with of these two variables.
direct measurements of the turbulent velocity. Of
the first aspect, I will focus on vertical distribu- Buoyancy and vertical velocity of eggs
tions and use fish eggs as examples, since their
physical properties are simple in a hydrodynamical The terminal velocity of a particle moving in a
sense and their biological properties are well fluid is the constant velocity reached when there is
described. The second aspect, encounter rates balance between gravity forces and frictional forces.
between planktonic particles, will be illustrated by For a sphere like a fish egg it is a function of the
investigations on how cod larvae feeding on cope- gravity acceleration, g, the egg diameter, d, the
podite nauplii are influenced by turbulence- buoyancy of the egg, ∆ρ= ρw - ρe, where ρe is the
induced particle interaction. specific gravity of the egg and ρw is the specific

160 S. SUNDBY
gravity of the ambient water. When the Reynolds’ found in the Stokes’ regime or in the lower part of the
number, Re= wd/ν (where ν is the molecular vis- intermediate regime, where the Reynolds’ number is
cosity of the water), is less than about 0.5, the vis- generally lower than 5. The value of KI in equation
cous forces dominate the frictional force and the ter- (4) is not a true constant in the whole range of the
minal velocity is expressed by Stokes equation (See intermediate region. However, for the actual range of
Tim Pedley´s lecture, this volume): Reynolds numbers, i.e. < 5, we assume that KI is
approximately constant. A numerical expression for
w = 1/18 g d2 ∆ρ ν-1 (3) KI can then be obtained by setting equations (3) and
(4) equal to the value of w at which Re=0.5. This
When the Reynolds’ number exceeds 0.5 viscous gives KI=19 (Sundby, 1983). In Figure 1 the terminal
forces are starting to become less important due to velocity is plotted against egg diameter for a range of
an increase of the turbulent forces. For a moderate- buoyancy values, ∆ρ, using equations (3) and (4).
ly large Reynolds’ number the frictional forces are After Coombs (1981) introduced the density-gra-
influenced by both viscous and turbulent friction dient column to measure the neutral buoyancy of
and an empirical expression for the terminal veloci- fish eggs, new possibilities appeared to understand
ty was found by Dallavalle (1948): the processes behind the vertical distribution of
ichthyoplankton. By this laboratory technique it is
w = KI (d -ζD) ∆ρ2/3 ν-1/3 (4) possible to measure the density of individual fish
eggs to an accuracy of about 0.04×10-3 g cm-3.
where KI is a constant, ζ is a constant equal to 0.4 Many species of fish eggs have a specific gravi-
for spheres and D is the uppermost limit of egg ty which makes them positively bouyant in the
diameter to which the Stokes equation applies. upper mixed layer. They have a neutral buoyancy,
Hence, the terminal velocity becomes linearly pro- measured as salinity, which is lower than the salini-
portional to the egg diameter. ty of the upper layer. They are distributed as group
For completely turbulent conditions, when the A in Figure 2 and are termed “pelagic” eggs (Sund-
Reynolds’ number reaches the order of 103, the ter- by, 1991). Four of the egg species plotted in Figure
minal velocity is proportional to the square root of the 1 are pelagic eggs. Some species have eggs which
diameter. It is shown in Figure 1 that fish eggs are are heavier than the upper layer, but lighter than the
density of the deep layer. They are distributed as
group B in Figure 2 and are termed “bathypelagic
eggs”. Examples of such types are eggs of Pacific

FIG. 1. – Terminal velocity as a function of egg diameter for a range FIG. 2. – Buoyancy distributions of three main groups of fish eggs in
of buoyancies, ∆ρ. Observed values of egg diameter and buoyan- relation to the salinity profile shown below, as defined by Sundby
cies in their natural habitats for five different species of fish eggs (1991). A. Pelagic eggs, ∆ρ > 0; B. Bathypelagic eggs, ∆ρ = 0; C.
are plotted on the graph. 1. North Sea mackerel; 2. Arcto-Norwe- Bottom eggs, ∆ρ < 0. The salinity profile shows a typical profile for
gian cod; 3. North Sea plaice; 4. Namibian Cape hake; 5. Atlantic coastal boreal regions with a upper mixed layer of lower salinity, a
halibut at the coast of northern Norway. halocline below and a deep homogeneous deep layer of high salinity.

TURBULENCE AND ICHTHYOPLANKTON 161


FIG. 3. – Section through an Atlantic cod egg with developed embryo.

halibut (Thompson and van Cleve, 1936), Baltic cod Arcto-Norwegian cod eggs it is quite small, only 3 -
(Kändler, 1949) and Atlantic halibut (Haug et al., 5% of the total egg volume.
1984). Group C in Figure 2 is defined as bottom The yolk and embryo constitute the light fraction
eggs and are heavier than the density of the bottom of the egg and is surrounded by the vitelline mem-
layer. Examples of such eggs are the saffron cod brane across which the osmoregulation occurs and
eggs in the northeast Pacific Ocean (Dunn and prevents the embryo and yolk from being dehydrat-
Matarese, 1986) and Barents Sea capelin eggs ed. The density of the embryo and yolk of Arcto-
(Bakke and Bjørke, 1973). It should, however, be Norwegian cod eggs is about 1.017 g cm-3. The
emphasized that eggs from various species may also ambient salinity of Arcto-Norwegian cod eggs is 33
have buoyancy distributions which fall between the - 34 p.s.u. In contrast, Baltic cod eggs, which devel-
three main groups defined above. Coombs et al. op in the halocline at low salinities, less than 10 -15
(1985), for example, showed that eggs of sprat and p.s.u., have developed eggs with a high water con-
pilchard off the south coast of Great Britain had a tent which makes them much more buoyant
buoyancy distribution between pelagic and bathy- (Nissling and Westin, 1991). This is an ecological
pelagic eggs and their buoyancy changed through adaptation in Baltic cod to prevent the eggs from
development. Kendall and Kim (1986) demonstrat- sinking down to the anoxic deep water of the Baltic
ed that bathypelagic eggs of walleye pollock may Sea. The specific gravity of the embryos of these
substantially change their vertical distribution due to eggs is about 1.008 g cm-3, and the eggs are big, 1.8
changes in buoyancy during the egg development. - 1.9 mm in diameter, and have a thin chorion. It has
We now explore the properties which determine been shown that the spawning female cod invests a
the buoyancy of eggs. I will use investigations on relatively constant mass of chorion in the eggs inde-
Atlantic cod eggs (Kjesbu et al., 1992) as an exam- pendent of egg size. This implies that bigger eggs
ple. Figure 3 shows a section through an Atlantic generally have thinner chorions. The specific gravi-
cod egg with the embryo partly developed. The ty of the egg, ρe , can then be expressed:
diameter of this egg ranges from 1.1 to 1.6 mm.
The egg shell, the chorion, is the heavy fraction of ρe = ρi + (ρch - ρi) Vch/(4/3πr3) (5)
the egg. Kjesbu et al. (1992) found that the specific
gravity of the chorion was 1.20 g . cm-3. The thick- where ρi is the specific gravity of the embryo and
ness is, however, only 5 - 9 microns. Holes in the yolk, ρch is the chorion specific gravity, Vch is the
chorion allows seawater to penetrate it, and sea chorion volume and r is the radius of the egg.
water of the same density as the ambient seawater is Figure 4 shows graphs of equation (5) for three
found in the perivitelline space, which is the space different values of chorion volumes, the maximum,
between the chorion and the vitelline membrane. the mean, and the minimum volumes. It can be seen
The volume of the perivitelline space may vary. For from the figure that egg size within the natural range

162 S. SUNDBY
rrent the vertical eddy diffusivity coefficient might
increase in and above the bottom boundary layer
due to bottom friction. Figure 5 shows qualitative-
ly how the vertical eddy diffusivity might vary
through a 100 m deep water column with the pyc-
nocline at 30 - 50 m depth.
Estimating the mixed layer eddy viscosity coeffi-
cient is difficult, partly due to technical problems in
measuring the wave zone. Sverdrup et al. (1942)
derived estimates of the eddy viscosity coefficient
from Ekman theory. Sundby (1983) estimated over-
all eddy diffusivity coefficients for the mixed layer
from a model based on the vertical distribution of
pelagic eggs. Thorpe (1984) estimated the eddy dif-
FIG. 4. – Egg specific gravity of Arcto-Norwegian cod as a function
of egg diameter for: the minimum chorion volume (26 x 106 mm3), fusivity coefficients in the surface layer based on a
lower line; the mean chorion volume (33 x 106 mm3), middle line; model of the vertical distribution of air bubbles in
and the maximum chorion volume (40 x 106 mm3), upper line.
the sea. Although their results differ to some extent,
it may be concluded that the eddy diffusivity coeffi-
of variation (1.1 - 1.6 mm) contribute more to vari- cient ranges from about 10 cm2 s-1 at wind speeds
ations in specific gravity than the variation in chori- near zero to about 103 cm2 s-1 during strong surface
on volume (26 x106 mm3 - 40 x106 mm3). mixing in stormy weather.
In the pycnocline the eddy diffusivity coeffi-
Vertical eddy diffusivity coefficient cient is inversely related to stratification and
directly dependent on energy input. Examples of
The other variable which influences the vertical authors who have estimated the vertical eddy dif-
distribution of eggs is the vertical eddy diffusivity fusivity in pycnoclines are Gade (1970) for
coefficient (K in equation (1) and (2)). Depending Oslofjorden, and Kullenberg (1971) for shallow
on depth, wind velocity, stratification, surface coastal waters. Gargett (1984) reviewed the litera-
cooling and convection, tidal energy and bottom ture on vertical diffusivity coefficients in stratified
stress, it varies over approximately five orders of systems. Depending on the level of stratification,
magnitude. It is normally largest in the mixed layer the eddy diffusivity coefficient ranged from
and decreases to a minimum in the pycnocline due 0.5x10-2 cm2 s-1 to 4 cm2 s-1.
to the strong buoyancy forces which act against Bottom turbulence, which normally extends sev-
vertical mixing. In regions of strong bottom cur- eral metres above the bottom, is mainly dependent
on the boundary layer velocity and bottom rough-
ness. Bowden (1962) reported values from several
authors. In areas of strong tidal mixing the eddy dif-
fusivity coefficient may exceed 100 cm2 s-1. Above
the seabed in deep oceanic areas 1 cm2 s-1 is more
common.

Analytical solution for pelagic eggs

Pelagic eggs which have a neutral buoyancy


distribution like A in Figure 2 will ascend towards
the surface. As we assume that there is no change
of density within the mixed layer, the vertical
velocity of eggs as a function of depth is constant
(w(z)=const.). If we also assume that the eddy dif-
FIG. 5. – Qualitative shape of the eddy diffusivity coefficient, K in fusivity coefficient through the mixed layer is con-
cm2s-1, (right part of the figure) for a hydrographic profile (left part
of the figure) identical with the profile in Figure 2. The shaded area stant (K(z) = const.), the solution to equation (2)
indicates typical range of values for eddy diffusivities. becomes:

TURBULENCE AND ICHTHYOPLANKTON 163


and diameter. However, as shown in Figure 2 one
population of eggs has a neutral buoyancy which is
most often Gaussian distributed around mean val-
ues. Sundby (1983) extended the solution to a
Gaussian distributed population of pelagic eggs.
This analytical solution is more complicated. How-
ever, for the buoyancy distributions in question
(Arcto-Norwegian cod eggs and North Sea macker-
el eggs), the results do not differ substantially from
the more simple solution of equation (6).
The extended solution by Sundby (1983) was
fitted by nonlinear regression to measured vertical
profiles of Arcto-Norwegian cod eggs off the coast
of Northern Norway for various wind situations,
FIG. 6. – Vertical distribution of pelagic eggs as shown in equa- and hence mixing conditions (Figure 7), and mea-
tion (6) for two values of w/K. Profile I: w/K is large. Profile II:
w/K is small. surements by Iversen (1973) of North Sea macker-
el eggs under wind conditions from Beaufort 0 to
6 (Figure 8).
The nonlinear regressions for the measurements
C(z) = C(a) exp - w/K (z-a) (6) of Arcto-Norwegian cod eggs, North Sea mackerel
eggs and North Sea Plaice eggs made it possible to
where C(a) is the egg concentration at a given depth, estimate the eddy diffusivity coefficients for the
a. Equation (6) is a simple exponential solution mixed layer, as a function of the wind speed (Figure
where the slope of the curve (Figure 6) is deter- 9) (Sundby 1983). The relation found was:
mined by the ratio w/K. When the ascending veloc-
ity of the eggs are large compared to the eddy diffu- K (cm2 s-1)= 76.1 + 2.26 W2 (7)
sivity coefficient (w/K is large) the egg concentra-
tion has the type I profile in Figure 6. When the where W is the wind speed in m s-1.
ascending velocity of the eggs is small compared to
the eddy diffusivity coefficient (w/K is small), the
egg concentration has the type II profile in Figure 6.
The solution in equation (6) is valid for one sin-
gle group of eggs with constant neutral buoyancy

FIG. 7. – Observed vertical distributions of Arcto-Norwegian cod FIG. 8. – Observed concentrations (crosses) of North Sea mackerel
eggs based on measurements of Solemdal and Sundby (1981) (dot- eggs by Iversen (1973) from Beaufort 0 to 6 wind force, and non-
ted lines) and nonlinear regression of the observed values based on linear regression of the observed values based on the solution by
the solution by Sundby (1983) (continuous lines). Sundby (1983).

164 S. SUNDBY
FIG. 9. – Computed eddy diffusivity coefficients, K, based on mea-
surements on vertical distributions of North Sea mackerel eggs,
North Sea plaice eggs, and Arcto-Norwegian cod eggs. FIG. 10. – Theoretical computed vertical profiles of egg concen-
(After Sundby 1983). tration of mackerel eggs and cod eggs at K=80 cm2 s-1 (wind speed
of 2 m s-1), K=400 cm2 s-1 (wind speed of 12 m s-1), and
K=1150 cm2 s-1 (wind speed of 22 m s-1).
The results are comparable with estimates by
other methods, and demonstrate that plankton, like conditions. During strong wind conditions (e.g. 22
fish eggs, can be applied to estimate physical prop- m s-1) vertical mixing dominates over the ascending
erties when precise measurements of buoyancy by velocity of the eggs and the difference between the
the method by Coombs (1981) are applied. In Figure vertical distribution of the two eggs species becomes
10, equation (7) is inserted in the model for pelagic negligible. The figure also shows that the egg pro-
eggs (equation (6)) to demonstrate how wind mixing files undergo the largest changes in the range
influences the vertical distribution of Arcto-Norwe- between 0 and 12 m s-1 of wind speed.
gian cod eggs and the lighter North Sea mackerel The analytical solution presented above is a
eggs with higher ascending velocity. As seen in the steady state solution. Westgård (1989) developed a
figure the lighter mackerel eggs are concentrated numerical solution which allows for studying tem-
more towards the surface during moderate wind poral variations. Figure 11 presents some applica-

FIG. 11. – Development of the vertical profile of pelagic eggs (type A buoyancy distribution as
shown in Figure 2) from an initial distribution of spawning near the bottom at 120 m depth to steady
state distribution. Two events of mixed layer turbulence are shown. Upper part: wind velocity,
W = 0 m s-1. Lower part : wind velocity , W = 15 m s-1. (After Sundby 1991).

TURBULENCE AND ICHTHYOPLANKTON 165


tion of this numerical model as shown by Sundby
(1991). The figure presents two scenarios of cod
spawning at 120 m depth in a salinity profile as
shown in Figures 2 and 5. The profiles display how
the eggs ascend from the spawning depth and final-
ly reach the steady state solution when there is bal-
ance between the buoyancy and eddy diffusion flux-
es. The upper profiles show that the equilibrium pro-
file is reached after about 48 hrs. during calm con-
ditions, i.e. without wind and wind-generated turbu-
lence. During strong wind-induced mixing (wind
speeds of 15 m s-1) the equilibrium profile is reached
after only 30 hrs.

Analytical solution for bathypelagic eggs FIG. 12. – Vertical distribution of bathypelagic eggs in a strong pyc-
nocline where the eddy diffusivity coefficient is small and in a weak
pycnocline with a larger eddy diffusivity coefficient.
Bahypelagic eggs which have a neutral buoy-
ancy distribution like B in Figure 3 will ascend
from the deep layer, but will descend from the σ = (K/m)1/2 (12)
upper mixed layer. Their neutral buoyancy is
found in the pycnocline. We consider one single around the the mean depth level zA.
buoyancy group of bathypelagic eggs. We assume When the velocity of the eggs is confined within
that the pycnocline is linear, and may therefore be the Stokes regime, the Stokes equation (equation 3)
expressed: for the terminal velocity is valid and the expression
for m in equation (12) becomes:
ρ(z) = kz + b (8)
m = 1/18 d2 ν-1 ρw N2 (13)
where ρ(z) is the density as a function of depth, z,
and k and b are constants. where N is the Brunt-Väisälä frequency (with the
The vertical velocity will then vary linearly with- unit s-1).
in the Stokes regime (equation 3), and may therefore Figure 12 shows the graphical form of the solu-
be written: tion (equation (11)). The distribution is narrow
when the eddy diffusivity coefficient is small and m
w(z) = m(z - zA) (9) (proportional to the the density gradient) is large.
To give an impression of the magnitude of the stan-
where m is a constant and zA is the depth level where dard deviation of the vertical distribution of bathy-
∆ρ(z) = 0, i.e. the level of neutral buoyancy of the pelagic eggs, let us take Atlantic halibut eggs in the
egg. We now assume that the eddy diffusivity coef- fjords of Northern Norway as an illustrative exam-
ficient is constant in a linear pycnocline, and equa- ple. Typically, as for the density profiles of Figure
tion (2) can then be written: 13, the Brunt-Väisälä frequency ranges from
0.5x10-4 to 2.0x10-4 s-2. From the above-mentioned
- m(z - zA) . C(z) = K dC(z)/dz (10) literature on the influence of stratification on the
turbulence, the eddy diffusivity coefficients range
The solution to equation (10) is: from 0.1 to 0.5 cm2 s-1. The diameter of the halibut
m eggs are large compared to other fish eggs, 3.0 - 3.5
C( z ) = C A exp ⎡⎢ − z − z A ) ⎤⎥
2
( (11) mm. When these values are inserted into equations
⎣ 2K ⎦ (12) and (13), the standard deviation, σ, of the ver-
where CA is the concentration of eggs at the depth of tical spreading of one bouyancy group of halibut
neutral buoyancy of the eggs, zA. It appears from eggs will range from 0.4 to 1.6 m. However,
equation (11) that the bathypelagic eggs are verti- according to Haug et al. (1986) (Figure 13), the
cally distributed as a normal distribution with a stan- older eggs (which have come to a steady state ver-
dard deviation: tical distribution) extend over a 150 - 250 m water

166 S. SUNDBY
FIG. 13. – Density profiles and profiles of Atlantic halibut eggs in fjords of Northern Norway (after
Haug et al. 1986).

column. Consequently, the large vertical spreading groups of eggs. Group I is the lightest fraction and
of halibut eggs observed in the water column must has a specific gravity of 1.0270 g cm-3 which give
be due to a spreading in the neutral buoyancy dis- eggs an equilibrium level at 120 m depth. Group II,
tribution of the eggs alone and is not caused by ver- the average fraction, has a specific gravity of
tical turbulence. 1.0272 g cm-3 with an equilibrium depth of 160 m.
Atlantic halibut eggs are spawned in the deep Group III, the heavy fraction, has a specific gravity
water below the pycnocline, most often near the bot- of 1.0274 g cm-3 with the equilibrium level at 200 m
tom at 400 - 800 m depth in Norwegian fjords. Con- depth. Despite the small differences in specific
sequently, they have to ascend several hundred gravity the time to reach the respective equilibrium
meters before they reach the level of equilibrium at depth levels are very different for the heavy group
the pycnocline. In Figure 14 one of the profiles from compared to the two other groups. Groups I and II
Haug et al. (1986) is used to demonstrate the time it reach the equilibrium level at about 4 - 5 days, while
takes to ascend from a spawning depth of 450 m. the heaviest fraction which has the equilibrium level
The right part of the figure shows vertical position at 200 m depth, only 250 m above the spawning
as a function of time in days for three buoyuancy depth will use more than 9 days to reach that level.

TURBULENCE AND ICHTHYOPLANKTON 167


alone by the moving predator. Neither prey nor the
fluid environment has a velocity. The number of
encountered prey per time unit is then linearly pro-
portional to the search volume of the predator, and
can be expressed by the relation:

Z = N π R2. v (14)

where Z is the contact rate between larvae and their


prey, N is the number of prey particles per volume
unit, R is the perceptive distance for the fish larva,
and v is the swimming speed of the fish larva.
FIG. 14. – Left panel: Atlantic Halibut eggs profile (dashed line) Hence, there is a linear relationship between the
and water column density profile (continuous line) (After Haug et
al. 1986); Mid panel: Calculated vertical velocities for three buoy- contact rate, Z, and the swimming speed, v. Gerrit-
ancy groups of eggs, i.e. respectively 1.0270, 1.0272 and 1.0274 g sen and Strickler (1977) pointed out that this is an
cm-3; Right panel: Calculated vertical positions as a function of time
for the three egg groups initially released at 450 m depth. oversimplification. They showed that the relative
velocity between predator and prey is not a function
of the predator velocity alone, but also of the prey
This is because the buoyancy of the heavy fraction velocity. They developed a model where the preda-
gives a very low ascending speed of those eggs in tor is cruising in a cloud of stochastically moving
the deep homogeneous layer. prey. When predator speed is higher than the prey
speed the contact rate becomes:

LARVAL FISH FEEDING AND THE ROLE OF Z = N π R2 (u2 + 3v2)/3v (15)


TURBULENCE-INDUCED CONTACT RATE
where u is the prey speed. The velocity component
Basic theory of the contact rate, (u2 + 3v2)/3v, in the equation
above is larger than the velocity component, v, of
The influence of turbulence on plankton equation (14). The middle panel of Figure 15
encounter rates is a biophysical process which sub- shows the effect of the Gerritsen and Strickler
stantially alters traditional considerations about (1977) model, and it implies that the fish larvae
energy demands and behaviour of larval fish, and will need to swim a shorter distance to maintain the
consequently the recruitment processes. The basic same encounter rate as in the situation of the upper
idea was presented by Rothschild and Osborn panel. Hence, the tube of swept volume between
(1988), and they developed the theoretical frame- each encountered prey is shorter than in the upper
work necessary to investigate the influence of turbu- panel.
lence on plankton contact rates in the laboratory and Let us use this model on the event of cod larvae
in the field. Their theory was in turn built on the the- feeding on copepod nauplii. The mean swimming
ory of Gerritsen and Strickler (1977) on the speed of these nauplii are only 10% of the mean
encounter rate between a cruising predator and a sto- cruising speed of the larvae. The enhancement of the
chastically moving prey. encounter rate using Gerritsen and Strickler’s model
To give an intuitive and qualitative understand- is (u2 + 3v2)/3, and for cod larvae cruising at mean
ing of the processes which enhance contact between speed it implies an enhancement in encounter rate of
predators and prey, let us look at Figure 15. The only 0.3%. Only when the cod larva is cruising very
three panels of the figure show a fish larva and its slowly, at speeds comparable to the speed of the
prey, for example copepod nauplii. The tubes of the nauplii, is there a considerable improvement in the
three panels conceptually indicate the swept volume contact rate, 33%, compared to the traditional model
by the fish larva between each prey encounter. The (equation (14)). Even though the Gerritsen and
upper panel indicates the traditional view of how a Strickler model did not give significant practical dif-
larval cruise predator searches through the water for ferences in encounter rate for the case of cod larvae,
prey. By this concept it is assumed that the relative it was principally a very important theoretical devel-
motion between the predator and prey is caused opment.

168 S. SUNDBY
FIG. 15. – Three model concepts of larval encounter rate. Upper panel: The traditional concept of
encounter based on swept area by the larva. Only the larva is moving. Middle panel: Encounter model
according to the Gerritsen and Strickler (1977) model with stochastically moving prey. Lower panel:
Encounter rate model according to Rothschild and Osborn (1988) with stochastically moving prey
and ambient turbulence.

Rothschild and Osborn (1988) pointed out that in the tube. The other aspect of this model is that the
nature turbulence contributes more to the stochastic natural turbulence becomes so important that the
velocity component than the prey. They added the encounter rate is in fact quite insensitive to the
turbulent velocity to the formulation in equation swimming speed of the cod larvae within its natural
(15) and arrived at a modified Gerritsen and Strick- range.
ler model: The nature of turbulence is that it occurs on all
scales at the same time and with a range of turbulent
Z = N π R2 (u2 + 3v2 + 4w2)/3(v2 + w2)1/2 (16) velocities increasing with increasing scale. There-
fore, we need to sort out how we can apply in equa-
where w is the root-mean-square turbulent velocity. tion (16) the range of turbulence velocities. The
For typical turbulent velocities in nature the basic relation of scales in natural turbulence (ocean-
velocity component in equation (16) becomes con- ic and atmospheric turbulence) as first postulated by
siderably larger than in equation (15), and the larva, Kolmogorov (1941), is that turbulent energy cas-
in the lower panel, will have to swim a considerably cades from large scale to smaller and smaller scales
shorter distance to maintain the same encounter rate until it dissipates to heat by shear strain of molecu-
as in the mid panel, again indicated by the length of lar movement. A universal relation between the

TURBULENCE AND ICHTHYOPLANKTON 169


energy spectrum and wave number, k, of turbulence Sundby and Fossum (1990) applied the theory of
of high Reynolds numbers in the inertial subrange Rothschild and Osborn (1988) to field observations
exists, and this is proportional to k-5/3. Therefore, tur- of first-feeding cod larvae feeding on Calanus fin-
bulence on all length scales are found simultaneous- marchicus nauplii. The turbulent scales which con-
ly in a body of water in the ocean. Many later mea- tribute to increase the contact rate are at all scales
surements from the field have confirmed such a tur- smaller than the typical separation distance between
bulent energy spectrum (e.g. Grant et al. 1962). The predator and prey. The typical nauplii concentra-
consequence of this physical nature is the well- tions, c, for first-feeding cod larvae are 1 - 50 liter-1,
known fact that eddy diffusivity increases with which correspond to deterministic separation dis-
increasing length scale (Okubo, 1978). This means tances, r=c-1/3, from 10 to 3 cm. Against this back-
that the spreading of a certain property in nature, for ground Sundby and Fossum (1990) used 5 cm as an
example fish eggs, increases with time and with the average separation distance. Depending on the level
size of the distribution. This is simply because larg- of turbulent kinetic energy dissipation, ε, the root-
er and larger turbulent eddies contribute to the mean-square velocity, w, for such separation dis-
spreading, or rearrangement of the patches of eggs. tances will typically range from the order of 0.01 -
Similarly, the relative motion between two individ- 1 cm s-1, as calculated according to Rothschild and
ual particles due to turbulence will increase as the Osborn (1988). By applying equation (4) of Roth-
distance between them increases. This basic process schild and Osborn (1988), valid for scales larger
contributes simultaneously to spreading of the parti- than the Kolmogorov scale, the relevant turbulent
cles and contact between them. So, both spreading velocity, w, can be calculated:
and contact rate is scale dependent in natural (aquat-
ic and atmospheric) turbulence. w = 1.9 (ε d)1/3 (17)
Rothschild and Osborn (1988) developed the
relation between the turbulent scales and the associ- where d is the turbulent length scale, and w is the
ated turbulent velocities, the root-mean-square turbulent velocity associated with that length scale.
velocity. This development was summarized in Fig- The turbulent length scales which increase the
ure 1 of their paper (Figure 16 here), and shows how contact rate between predator and prey are all length
the turbulent velocity increases with increasing scales smaller than the separation distance between
length scale. the predator and prey, because all such length scales
contribute to the relative velocity between them. For
fish larvae where the concentration is typically
much lower than the concentration of the prey items,
the separation distance is effectively equal to the
separation distance between the prey items, r. Tur-
bulent cells of a scale, d, larger than the separation
distance, r, will not contribute to increase the contact
rate, because those cells will move larger water
parcels without rearranging the relative distance
between the particles.
To get the right impression of the proper length
scales involved in plankton contact rates, and partic-
ularly the length scales involved in feeding cod lar-
vae, let us look at Figures 17 - 20. Figure 17 shows
the main prey organism of Arcto-Norwegian cod
larvae, a Calanus finmarchicus nauplius stage III.
The dots around the nauplius indicate the average
distance between phytoplankters in a typical spring
bloom of 106 cell l-1. In Figure 18 a first-feeding cod
larva is shown with a Calanus finmarchicus nau-
FIG. 16. – The relation between length scale, separation distance, plius stage III, at the maximum reactive perceptive
and the associated root-mean-square turbulent velocity at various
levels of turbulent energy dissipation rates (from Rothschild and distance which is about 0.45 cm. The vectors indi-
Osborn, 1988). cate the average swimming speeds of the larva and

170 S. SUNDBY
FIG. 17. – Calanus finmarchicus nauplius stage III, the main prey
organism for first-feeding Arcto-Norwegian cod larvae. The dots
indicate the food concentration of phytoplankters in a normal
spring bloom.

FIG. 19. – Cod larva (with the maximum perceptive distance indi-
cated by the half circle) in a very high concentration of nauplii, 50
nauplii l-1. The dimensional size of the nauplii are exaggerated; oth-
erwise they would have been invisible. The square shaped clouds of
dots around two of the nauplii indicate the concentration of a phy-
toplankton spring bloom of 106 plankters l-1 and a concentration of
105 plankters l-1. Also the dimensional size of the
phytoplankters are exaggerated to allow visibility.

FIG. 18. – First-feeding cod larva and its main prey, Calanus fin-
marchicus nauplius stage III. The circle indicates the maximum
reactive perceptive distance, 0.5 cm, and the arrows indicate the
mean swimming speeds of the two organisms, 0.17 and 0.02 cm s-1,
respectively.
FIG. 20. – Cod larva (with the maximum perceptive distance indi-
cated by the half circle) in a low concentration of nauplii, 1 nauplii
l-1, but still high enough concentration to survive in turbulent con-
the nauplius, 0.17 and 0.02 cm s-1, respectively. Fig- ditions. The dimensional size of the nauplii are exaggerated;
ure 19 shows the cod larva swimming in a situation otherwise they would have been invisible.
of very high food concentration of Calanus nauplii,
50 nauplii litre-1. The square-shaped clouds of parti- Application of theory on field data
cles around two of the nauplii indicate the phyto-
plankton particle densities in a spring bloom of 106 Field evidence for the influence of turbulence on
cells per litre, which is the same concentration as the feeding rate of fish larvae has been provided by
shown in Figure 17, and of a particle density of 105 Sundby and Fossum (1990) and Sundby et al.
cell l-1. Figure 20 shows the first-feeding cod larva (1994). They investigated the gut content, i.e. the
in a low concentration of prey, 1 nauplii l-1. Even average number of nauplii in the gut, of cod larvae
this low concentration has been shown to give suffi- sampled at the first-feeding areas in Lofoten, North-
cient feeding conditions for cod larvae under turbu- ern Norway, and compared it to the nauplii concen-
lent conditions (Sundby and Fossum, 1990). tration in the sea under various wind situations.

TURBULENCE AND ICHTHYOPLANKTON 171


The relation between the average number of prey
particles in the larval gut, A(c), (termed feeding
ratio) and the nauplii concentration in the sea, c, can
be expressed as follows:

A(c) = Amax [1 - e-(b · c)] (18)

where Amax is the maximum number of prey in the


larval gut, i.e. full gut, and b is the coefficient
which determines the slope of the function in equa-
tion (18). The function is shown in Figure 21. The
coefficient, b, contains all the environmental fac-
tors which influence larval feeding. Light condi-
tions and turbulence are the most important factors FIG. 22. – Feeding ratio of cod larvae versus naupliar concentration
influencing larval feeding in addition to prey con- in the sea for all data (259 points) presented by Sundby et al.
(1994). One point represents one sample depth. Graphs of equation
centration. (18) for b = 0.02 and 0.2 are plotted into the figure.
Sundby and Fossum (1990) and Sundby et al.
(1994) fitted their field data to equation (18) by
nonlinear regression analysis. Data sampled dur- feeding. The panels in Figure 23 show larval feed-
ing night time (10 p.m. to 2 a.m.) were omitted in ing under increasing wind conditions from wind
order to minimize the effect of reduced light on the speeds of 2.0 m s-1 to 10.5 m s-1, grouped in four
feeding conditions. During this period the feeding wind speed intervals: 2.0 m s-1, 3.2 - 4.0 m s-1, 5.4 -
ratio is lower than during day time. By omitting 6.9 m s-1, and 7.5 - 10.5 m s-1. In all there are 13
the data from the dark part of the day, it is expect- larval sampling stations distributed in the four
ed that the variations in the coefficient b are main- wind speed intervals. The different values of the
ly caused by variations in the turbulence-induced coefficient, b, for the 13 larval sampling stations
feeding rate. In Figure 22 all the data sampled by are plotted against wind speed in Figure 24. The
Sundby et al. (1994) is presented. Two graphs of figure indicates that turbulence-induced feeding
equation (18) for b = 0.02 and b = 0.2 are plotted rate is increased by a factor of 9 when the wind
onto the figure. Initially it seems that the data indi- speed increases from 2 to 10 m s-1.
cate no functional relationship. However, when
the data sampled under equal turbulent conditions
with respect to wind-induced turbulence were
grouped together and the coefficient b was deter-
mined for each event (Figure 23), it is evident that
wind-induced turbulence together with food con-
centration are the main causes of variable larval

FIG. 23. – Feeding ratio of cod larvae versus naupliar concentration


FIG. 21. – The functional relation between number of nauplii in the in the sea. The data are pooled into four wind groups. Panel 1: 2.0
larval gut (here termed feeding ratio) and nauplii concentration in m s-1; Panel 2: 3.2 - 4.0 m s-1; Panel 3: 5.4 - 6.9 m s-1; Panel 4:
the sea. 7.5 - 10.5 m s-1.

172 S. SUNDBY
FIG. 24. – The coefficient b in equation (18), derived from nonlinear FIG. 25. – Theorically calculated contact rate (after the Roth-
regression, as a function of wind speed. schild and Osborn formulation) as a function of larval swimming
speed for four different turbulent conditions. Lower line is with-
out turbulence. The upper lines are for wind speeds of 2.5, 5.0,
and 7.5 m s-1. The dotted region shows the observed range of cod
At a wind speed of 4 m s-1 the turbulent veloci- larval swimming speeds.
ty contributes more to contact rate than the larval
swimming speed. Figure 25 shows the contact rate the field is lower than in the laboratory where tur-
as a function of swimming speed for 4 different bulence intensities most often are low.
events of turbulent conditions. The lower graph Above I have considered feeding by first-feeding
shows the relation in the absense of turbulence, i.e fish larvae which are small and slow-swimming
the Gerritsen and Strickler (1977) formulation organisms compared to the turbulence velocities on
(equation (15)). The upper three graphs show the that scale. However, larger-scale turbulence may
Rothschild and Osborn (1988) formulation for tur- also have the potential to enhance contact rates for
bulent conditions corresponding to wind speeds of larger organisms, if the separation distance between
2.5, 5.0 and 7.5 m s-1 respectively. The ranges of their prey is comparably larger. Sundby (1995)
the observed swimming speeds of first-feeding cod analysed the influence of wind-induced turbulence
larvae in the laboratory (Solberg and Tilset, 1984) on cod larvae as they grow, increase their swimming
are indicated by the shading. It is clear from the speed and change to larger prey items of lower con-
figure that at wind speeds above 5 m s-1 the benefit centrations.
for the larvae to increase the contact rate by its own Based on data of post-larval-cod growth rate
motion is negligible, and the larvae will, conse- Sundby (1995) arrived at a relation for the average
quently, waste its energy resources if cruising for swimming speed as a function of age:
food, because turbulence will do the job anyway.
From this point of view the larvae should save its v(t) = 0.152.exp{5.9.[1-exp(-0.0050 t)]} (19)
energy resources for the short final attack on the
prey and for rapid escape reactions from predators. where v(t) is the age dependent swimming speed
Munk and Kiørboe (1985) indeed showed that this and t is age in days.
is what occurs with herring larvae when the As the Arcto-Norwegian cod larvae grow they
encounter rate increases: The swimming activity is change mainly to feed on larger copepodite stages
reduced and the attact rate increases. Larval growth of Calanus finmarchicus (Sysoeva and Degtereva,
itself is an important factor for survival, and it is 1965). The main diet of early juvenile cod (70 d)
unlikely that the larvae are adaptated to a energy- consists of copepodite stages IV and V (Helle,
wasting behaviour like excessive cruising. Turbu- 1994).
lence (together with light) is the most characteris- Data on copepodite concentration from the Nor-
tic feature of the mixed layer, the site where most wegian Sea and the coast of Norway were assem-
of the plankton production occurs, and therefore it bled from the literature. These data are synthesized
is reasonable to believe that the cruising speed in in equation (20):

TURBULENCE AND ICHTHYOPLANKTON 173


N(t) = No e- (0.03 · t) (20)

where N(t) is the age-dependent (in days) cope-


podite concentration. No is the concentration of first-
feeding prey (nauplii stage III). These data were
used to calculate the prey separation distance as a
function of time.
Figure 26 shows the effect of wind-induced mix-
ing on the contact rate between the cod and the cope-
pod nauplii/copepodites, as the cod grows from larva
to juvenile. It shows the relative increase in contact
rate from first-feeding larvae to early juveniles for
three different wind speeds: 5 m s-1, 10 m s-1 and 15
m s-1, and for three different prey concentrations
(low, average, and high). The three prey concentra-
tions correspond to the low, average, and high con-
centration of equation (20). The lines converge
toward 1 when the turbulent velocity becomes much
less than the swimming speed of cod, and conse- FIG. 26. – The relative increase (compared to non-turbulent condi-
tions) in the velocity component of the contact rate between Arcto-
quently, no longer contributes to the contact rate Norwegian cod and its prey from first-feeding stage to 65 d after
between the cod and its prey. The turbulence induced hatching, for three wind speeds, 5 m s-1, 10 m s-1, and 15 m s-1, and
for three prey concentrations according to equation 5. Hatched
contact rate is at its maximum for the slow-moving, lines: Nlow = 0.8 e -0.03 t . Thick lines: Nav. = 6 e -0.03 t. Dotted lines:
first-feeding cod larvae. The contact rate increases Nhigh = 50 e -0.03 t .
by an order of eight for the situation of low prey con-
centration under a wind speed of 15 m s-1 compared
to non-turbulent conditions, while it increases by a CONCLUDING REMARKS
factor of five for high prey concentrations under the
same wind speed. Hence, the turbulence induced Previous opinions that turbulence generally has
contact rate contributes to reduce the effects of vari- detrimental effects on planktonic life is in opposi-
able prey concentrations, since the ratio between the tion to theories and measurements developed over
maximum and minimum prey concentration, the past two decades. The mixed layer is the larval
Nhigh/Nlow, is 63, while the change in concentration habitat for the major proportion of fish species. As
experienced by the cod larvae (i.e. the change in the mentioned in the Introduction, light and turbulence
number of encounters) is less: Zhigh/Zlow = 39. This is are the most predominant energy sources in this
due to the fact that at lower prey concentrations the part of the ocean. Organisms are likely to have
separation between prey particles becomes larger developed strategies to benefit from the effects of
and, consequently, larger turbulent cells with higher these energy sources. MacKenzie et al. (1994)
kinetic energy contribute to increase the contact rate. modeled the optimal level of wind-induced turbu-
As the cod larvae grow and increase their swim- lence with respect to cod larvae encountering prey.
ming speed, the influence of the turbulence-induced They found that the wind speed giving optimal tur-
contact rate decreases. However, the simulations bulence is probably at the level of about 15 m s-1.
show that for strong mixing events the enhancement The field results from Sundby et al. (1994) showed
of contact rate is still considerable for 2-mo-old that the optimal level must be higher than wind
juveniles. At wind speeds of 15 m s-1 and at average speed of 10 m s-1, which is not inconsistent with
prey concentrations, the contact rate is higher by a MacKenzie et al. (1994). This implies that the fish
factor of 2.5 than it would be in non-turbulent con- larvae are able to benefit from quite strong turbu-
ditions. The prey concentration decreases as a func- lent mixing. In the first-feeding areas of cod larvae
tion of time, and again, as mentioned in the para- on the coast of northern Norway, the average
graph above, the increasing separation distance monthly wind speed (in May) is 5.6 m s-1 (Sundby,
between the prey particles contributes to maintain 1982), and wind speeds of 15 m s-1 or higher occur
the influence of turbulence as larger and larger less than 3% of the time. This leads to the conclu-
scales of turbulence come into effect. sion that detrimental effects of wind-induced tur-

174 S. SUNDBY
bulence for pelagic larval cod encountering prey REFERENCES
rarely exist.
The models for vertical distribution of eggs show Alcaraz, M., E. Saiz, C. Marrasé and D. Vaqué. – 1988. Effects of
turbulence on the development of phytoplankton biomass and
that it is important to distinguish between stationary copepod populations in marine microcosmos. Mar. Ecol. Prog.
distributions, where balance occurs between the Ser., 49: 117-125.
Bakke, S. and H. Bjorke. – 1973. Diving observations on Barents
buoyancy forces and the vertical turbulent mixing, sea capelin at the spawning grounds off northen Norway. Fisk.
and non-stationary distributions which are typical Dir. Skr. Ser. HavUnders., 16: 140-147.
Bowden, K.F. – 1962. Turbulence. In: M.W. Hill (ed.): The Sea.
for newly spawned eggs of less than 1 - 5 days old. Ideas and observations on progress in the study of the seas. Vol.
However, it is also important to realise that even a 1. Physical Oceanography, pp. 802-825. Interscience Publish-
ers. John Wiley & Sons.
stationary vertical distribution is not a static distrib- Coombs, S. H. – 1981. A density-gradient column for determining
ution on the individual particle level. The eggs are the specific gravity of fish eggs with particular reference to eggs
of the mackerel (Scomber scombrus). Mar. Biol . 63: 101-106.
continuously changing positions in relation to each Coombs, S.H., C.A. Fosh and M.A. Keen. – 1985. The buoyancy
other at a rate depending on the level of turbulent and vertical distribution of eggs of sprat (Sprattus sprattus) and
pilchard (Sardina pilchardus). J. Mar. Biol. Assoc. of the Unit-
energy dissipation. For larval distributions, individ- ed Kingdom , 65: 461-474.
ual behaviour adds to the change of vertical posi- Cury, P. and C. Roy. – 1989. Optimal environmental window and
pelagic fish recruitment success in upwelling areas. Can. J.
tion. Sclafani et al. (1993) showed that the stochas- Fish. aquat. Sci., 46: 670-680.
tically vertical movement of larvae may result in a Dallavalle, J.M. – 1948. Micromeritics. The technology of fine par-
ticles. Pitman, New York.
sensed prey concentration for the individual larvae Dunn, J.R. and A.C. Matarese. – 1986. A review of the early life his-
torheast Pacific Gadoid fishes. In: M. Alton (compiler): Work-
which deviates considerably from the average lar- shop on comparative biology, assessment, and management of
vae. When there are large variations in the vertical gadoids from the Pacific and Atlantic Oceans. Seattle, 24-28
June 1985. pp. 589-628. Northwest and Alaska Fisheries C.
distribution of predator and prey, and particularly, Gade, H.G. – 1970. Hydrographic investigation in the Oslo fjord, a
when peak concentrations do not overlap, the real study of water circulation and exchange processes. Geophysical
Institute, University of Bergen, Norway. Report no. 24.
encounter rate between predator and prey will be Gargett, A.E. – 1984. Vertical eddy diffusivity in the ocean interi-
complicated to predict. or. J. Mar. Res., 42: 359-393.
Gerritsen, J. and J.R. Strickler. – 1977. Encounter probabilities and
The topic of relevant turbulent scales for plank- community structure in zooplankton: a mathematical model. J.
ton contact rates presently lacks a mathematical for- Fish. Res. Bd. Can., 34: 73-82.
Grant, H.L., R.W. Stewart and A. Moilliet. – 1962. Turbulence
mulation. In this presentation I have from physical spectra from a tidal channel. J Fluid Mech., 12: 241-263.
reasoning concluded that all turbulent “cells” of Haug, T., E. Kjørsvik and P. Solemdal. – 1984. Vertical distribution
of Atlantic halibut (Hippoglossus hippoglossus) eggs. Can. J.
length scales less than the separation distance Fish. aquat. Sci., 41: 798-804.
between predator and prey contribute to increase the Haug, T., E. Kjørsvik and P. Solemdal. – 1986. Influence of some
physical and biological factors on the density and vertical dis-
contact rate. This is because turbulent motion on tribution of Atlantic halibut (Hippoglossus hippoglossus) eggs.
these length scales will continuously change the rel- Mar.Ecol. Prog. Ser., 33: 207-216.
Helle, K.– 1994. Size distribution of early juvenile Arcto-Norwe-
ative position between predator and prey. In the cal- gian cod in relation to food abundance and water mass prop-
culations above based on larval cod from the field I erties. Cand. scient thesis. University of Bergen.
Iversen, S.A. – 1973. Utbredelse og mengde av makrell egg
have, as an approximation, used an average concen- (Scomber scombrus) og zooplankton i Skagerrak og nordlige
tration of prey (i.e. Calanus finmarchicus nauplii) as delen av Nordsjøen i årene (1968-1972). Can. neal. thesis. Uni-
versity of Bergen.
the basis for calculating the separation distance and, Kändler, R. – 1949. Untersuchungen über den Ostseedorsch wärend
hence, the relevant turbulent scale. Rothschild der Forschungsfahrten mit dem R.F.D. «Poseidon» in den
Jahren 1925-1938. Berichte der Deutschen Wissenschaftlichen
(1988) stated that it is important to distinguish Kommision für Meeresforschung. Neue Folge , 11: 162-168.
between the deterministic separation distance, r, and Kendall, A.W., Jr. and S. Kim. – 1986. Buoyancy of walleye pol-
lock (Theragra chalcogramma) eggs in relation to water prop-
the probabilistic separation distance which is 0.55.r. erties and movement in the Shelikof Strait, Gulf of Alaska.
According to equation (17), the turbulence velocity Can. Spec. Publ. Fish aquat. Sci., 108: 169-180.
Kjesbu, O.S., H. Kryvi, S. Sundby and P. Solemdal. – 1992. Buoy-
associated with the probabilistic separation distance ancy variations in eggs of Atlantic cod (Gadus morhua L.) in
is 18% less than the turbulence velocity associated relation to chorion thickness and egg size: theory and observa-
tions. J. Fish Biol. 41: 581-599.
with the deterministic separation distance. The level Kolmogorov, A.N. – 1941. The local structure of turbulence in an
of prey patchiness will influence whether the prob- incompressible viscous fluid for very large Reynolds numbers.
C.R. Acad. Sci., USSR, 30: 301-305.
abilistic or deterministic distance is most relevant in Kullenberg, G. – 1971. Vertical diffusion in shallow waters. Tellus,
this context. However, the difference in the turbulent 23: 129-135.
Lasker, R. – 1975. Field criteria for survival of anchovy larvae: the
velocities associated with these two scales is rather relation between inshore clorophyll maximum layers and suc-
small compared to the uncertainties for many of the cessful first feeding. Fish. Bull. U.S., 73: 453-462.
MacKenzie, B.R., T.J. Miller, S. Cyr and W.C. Leggett. – 1994. Evi-
behavioural variables linked to the problem of dence for a dome-shaped relationship between turbulence and lar-
encounter rate between larvae and their prey. val fish ingestion rates. Limnol. Oceanogr. 39: 1790-1799.

TURBULENCE AND ICHTHYOPLANKTON 175


Mann, K.H. and J.R.N. Lazier. – 1991. Dynamics of marine ecosys- Sundby, S. – 1982. Investigations in Vestfjorden 1978. 1. Fresh
tems: biological-physical interactions in the oceans. Blackwell water budget and wind conditions. Fisk. Havet, 1978: pp.1-30.
Scientific Publications, Inc., Boston, USA. (in Norwegian, English abstract).
Munk, P. and T. Kiørboe. – 1985. Feeding behaviour and swim- Sundby, S. – 1983. A one-dimensional model for the vertical distri-
ming activity of larval herring (Clupea harengus) in relation to bution of pelagic fish eggs in the mixed layer. Deep-Sea Res.,
density of copepod nauplii. Mar. Ecol. Prog. Ser., 24: 15-21. 30(6A): 645-661.
Nissling, A. and L. Westin. – 1991. Egg buoyancy of Baltic cod Sundby, S. and P. Fossum. – 1990. Feeding conditions of Arcto-
(Gadus morhua) and its implications for cod stock fluctuations Norwegian cod larvae compared to the Rothschild-Osborn the-
in the Baltic. Mar. Biol, 111: 33-35. ory on small-scale turbulence and plankton contact rates. J.
Okubo, A. – 1978. Horizontal dipersion and critical scales for phy- Plankton Res., 12: 1153-1162.
toplankton patches. In: J.H. Steele (ed.) Spatial pattern in Sundby, S. – 1991. Factors affecting the vertical distribution of
plankton communities. pp. 21-42. Plenum Press, New York eggs. ICES mar. Sci. Symp., 192: 33-38.
Rothschild, B.J. – 1988. Biodynamics of the sea: the ecology of Sundby, S., B. Ellertsen, and P. Fossum. – 1994. Encounter rates
high dimensionality systems. In: B.J. Rothschild (ed.): between first-feeding cod larvae and their prey during moderate
Toward a Theory on Biological-Physical Interactions in the to strong turbulent mixing. ICES marine Science Symposiums,
World Ocean. NATO ASI Series. Series C: Mathematical and 198: 393-405.
Physical Sciences - Vol. 239, pp. 527-548, Kluwer Academic Sundby, S. – 1995. Wind climate and foraging of larval and juve-
Publishers. nile Arcto-Norwegian cod. In: R.J. Beamish (ed.): Climate
Rothschild, B.J. and T.R. Osborn. – 1988. Small-scale turbulence change and northern fish populations. Can. Spec. Publ. Fish.
and plankton contact rates. J. Plankton Res.,10: 465-474. Aquat. Sci., 121: 405-415.
Saiz, E. and M. Alcaraz. – 1992. Free-swimming behaviour of Sverdrup, H.U., M.W. Johson and R.H. Fleming. – 1942. The
Acartia clausi (Copepoda: Calanoida) under turbulent water oceans, their physics, chemistry and general biology. pp. 481-
movement. Mar. Ecol. Prog. Ser., 80: 229-236. 484. Prentice Hall, New York.
Sclafani, M., C.T. Taggart and K.R. Thompson. – 1993. Condi- Sysoeva, T.K. and A.A. Degtereva. – 1965. The relation between
tion, buoyancy and distribution of larval fish: implications the feeding of cod larvae and pelagic fry and the distribution
for the vertical migration and retention. J. Plankton Res., 15: and abundance of their principal food organisms. ICNAF Spe-
413-435. cial Publication, 6: 397-410.
Solberg, T. and S. Tilset. – 1984. Growth, energy consumption and Thompson, W.F. and R. Van Cleve. – 1936. Life history of the
prey density requirements in the first feeding larvae of cod Pacific halibut. (2) Distribution and early life history. Rep. Int.
(Gadus morhua L.) In: E. Dahl, D.S. Danielssen, E. Moksness, Fish. Comm., 9: 1-184.
and P. Solemdal (eds.): The propagation of cod (Gadus morhua Thorpe, S.A. – 1984. A model of the turbulent diffusion of bubbles
L.), Flødevigen rapportser., 1, 1984. below the sea surface. J. of Physical Oceanography, 14: 841-854.
Solemdal, P. and S. Sundby. – 1981. Vertical distribution of pelag- Westgård. T. – 1989. A model of the vertical distribution of pelag-
ic fish eggs in relation to species, spawning behaviour and wind ic fish eggs in the turbulent upper layers of the ocean. Rapports
conditions. Council Meeting of the International Council for the et Procès Verbaux des Réunions - Conseil International pour
Exploration of the Sea, 1981/77. l´Exploration de la Mer, 191: 195-200.

176 S. SUNDBY

You might also like