978 0883850305
978 0883850305
978 0883850305
DAVID WILDISH
Biological Station, St. Andrews, Canada
DAVID KRISTMANSON
University of New Brunswick
CAMBRIDGE
UNIVERSITY PRESS
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo
Published in the United States of America by Cambridge University Press, New York
Preface page ix
Acknowledgments xi
Introduction 1
Asking the right question 3
Scientific methods 7
Scientific communication 9
Aims and structure 12
References 15
Methods of study 17
Flow simulation 18
Velocity measurement 30
Seston sampling 38
Specialized methods 43
References 60
Dispersal and settlement 69
Benthic suspension feeder life cycle 69
Dispersal 73
Larval dispersal 74
Post-larval dispersal 91
Local hydrodynamics 93
Larval settlement/recruitment 95
Summary 117
References 120
Flow and the physiology of filtration 128
Background 128
Passive suspension feeders 132
Active suspension feeders 138
vn
Vlll Contents
Glossary 391
Index 405
Preface
IX
x Preface
XI
xii Acknowledgments
1
Benthic Suspension Feeders and Flow
Size range
Name (urn) Examples
Table 1.2. Some general questions related to the disciplines listed and
referable to suspension feeders or the communities in which they live.
Type Description
Observation Simple: involves direct experience, e.g. looking down a
microscope. Includes descriptive taxonomy, natural
history observations.
Complex: sampling or measurement which may involve
bias, e.g. benthic grab sampling involving different sieve
mesh sizes in sorting the animals.
Theory/model Reductionist: a way of looking at a subfield of benthic
biology which can be represented verbally or by a simple
conceptual model, e.g. the benthic limitation by flow
theory (see Chap. 8).
Holistic: as above, except that because of complexity or
the nature of the model, formal mathematical
representation is required, e.g. an ecosystem simulation
of a bivalve reef based on energy flow.
Inferential Laboratory or field experimental tests of formulated null,
experimentation Ho, and alternate, H1? hypotheses.
Criticism Logical refutation of any conclusions or constructs
derived from any of the types of scientific method
described here.
Scientific methods
Science is concerned with the creation and communication of knowl-
edge. Concerning the first of these, scientific knowledge is created by one
of the four major methods shown in Table 1.4.
Observation, in both the field and laboratory, is fundamentally impor-
tant to the scientific enterprise. During this process, realistic questions
and, hence, theories, models, or hypotheses are formulated. Such formu-
lations are possible constructs determined by logical reasoning concern-
ing the relations of benthic animals to the observed external world.
As pointed out by Peters (1991), there is no consensus on how words
like theory, model, and hypothesis are used, and he recommends using
them synonymously. In our account, we define hypothesis as a working
explanation of observations proposed in such a way that an experimental
test can choose between two contrasting constructs - null and alternate.
The hypothesis may apply fully, or only partially, to the theory or model
constructs which are conceived to be at a higher hierarchical level. The
Benthic Suspension Feeders and Flow
Scientific communication
Concerning the communication of science results mentioned earlier, the
active participants in science tend to form into natural groups which are
referred to as an "invisible college" because of shared research interests
and problems. Communication is the most important function of the
invisible college and various ways have been tried to optimize it. These
include personal communication (often electronic), workshops, sympo-
sia, published periodicals, and books.
In regard to periodicals, there are four distinct types: science maga-
zines, which provide interpretive articles for a multidisciplinary audi-
ence, but not original research articles (e.g. New Scientist, Scientific
American); technology transfer periodicals, whose purpose is to transfer
science to a commercially active field such as aquaculture (e.g. World
Aquaculture) but which does not usually carry original research; a few
multidisciplinary journals, notably Science and Nature, which contain
10 Benthic Suspension Feeders and Flow
Chapters
lOldl iVdUK
Journal 2 3 4 5 6 7 8 citations order
J. Exp. Mar. Biol. Ecol. 16 17 21 4 16 13 15 102 1
Mar. Ecol. Prog. Ser. 11 14 13 8 7 13 14 80 2
Mar. Biol. 15 10 14 11 13 7 8 78 3
Limnol. Oceanogr. 24 15 1 2 6 8 7 63 4
Biol. Bull. 7 1 7 6 5 0 1 27 5
Can. J. Fish. Aquat. Sci. 7 3 1 0 5 7 4 27 6
J. Mar. Res. 5 6 1 0 3 5 4 24 7
J. Mar. Biol. Assoc. U.K. 6 4 5 1 2 2 0 20 8
Science 0 4 3 3 4 1 5 20 9
Can. J. Zool. 2 3 0 2 6 0 0 13 10
Helgol. Meersunters. 4 1 1 0 1 3 1 11 11
Ophelia 3 2 2 1 0 1 1 10 12
Oceanogr. Mar. Biol. Ann. Rev. 1 4 0 1 0 3 0 9 13
Nature 3 3 0 0 2 0 1 9 14
Bull. Mar. Sci. 1 7 0 0 0 0 0 8 15
Neth. J. Sea Res. 2 2 0 0 0 2 1 7 16
Other journals 36 31 11 16 36 25 41 196
Books, Proceedings, etc. 18 12 9 11 19 13 13 95
Totals 161 139 89 66 125 103 116 799
total of 196 citations here represents 25% of the overall number of cited
articles.
a
Note: Full citations are given at the end of the appropriate chapter.
(Table 1.6), did not begin to be adequately addressed until the 1970s,
except by Crisp (1953). Thus, most of the research that will be discussed
is less than 25 years old.
The decision to limit the questions considered to proximate ones we
believe to be justifiable and liberating. It is liberating because it is then
possible to use the same experimental approach of strong inference
introduced by Francis Bacon in 1620, rounded out by the method of
multiple hypotheses advocated by Platt (1964), for most of the studies
reviewed herein. Purely descriptive or theoretical/model studies, unless
they have been tested by inferential experimentation, have been omitted
or de-emphasized. The exclusion of evolutionary questions is also liber-
ating because it absolves us from the responsibility to devise satisfactory
methods of study. That is not to say that benthic biologists should not be
interested in questions involving neo-Darwinian evolution (see Chap. 9).
Another difficulty we faced during the preparation of the book was,
Whom are we writing it for, and what should be its style? We view our
audience as primarily active research scientists, either already profes-
14 Benthic Suspension Feeders and Flow
References
Alcock, J. 1989. Animal behaviour: an evolutionary approach. 4th ed.
Sinnauer Associates, Sunderland, Mass. p. 596.
Chamberlin, T. C. 1965. The method of multiple working hypotheses. Science
148: 754-759.
Cherrett, J. M. 1989. Key concepts: the results of a survey of our members'
opinions, p. 1-16. In J. M. Cherrett (ed.) Ecological concepts: the
contribution of ecology to an understanding of the natural world.
Blackwell Scientific, Oxford.
Crisp, D. J. 1953. Changes in the orientation of barnacles of certain species in
relation to water currents. J. Anim. Ecol. 22: 331-343.
Darwin, C. 1842. The structure and distribution of coral reefs. Smith, Elder,
London.
Daugherty, R. L., J. B. Franzini, and E. J. Finnemore. 1985. Fluid mechanics
with engineering applications. 8th ed. McGraw-Hill, New York.
Denny, M. W. 1988. Biology and the mechanics of the wave swept
environment. Princeton University Press, Princeton, N. J.
1993. Air and water: the biology and physics of life's media. Princeton
University Press, Princeton, N. J.
Fenchel, T. 1987. Ecology - potentials and limitations. Ecology Institute,
Oldendorf/Luhe.
Hildrew, A. G., and P. S. Giller. 1994. Patchiness, species interactions and
disturbance in the stream benthos, p. 21-62. In P. S. Giller, A. G.
Hildrew, and D. F. Raffaelli (eds.) Aquatic ecology, scale, pattern and
process. Blackwell, Oxford.
J0rgensen, C B. 1966. Biology of suspension feeding. Pergamon Press,
London.
1990. Bivalve filter feeding: hydrodynamics, bioenergetics, physiology and
ecology. Olsen and Olsen, Fredensborg, Denmark.
Lincoln, R. J., G. A. Boxshall, and P. F. Clark. 1982. A dictionary of ecology,
evolution and systematics. Cambridge University Press, Cambridge.
Maddox, J. 1992. Electronic journals have a future. Nature 356: 559.
Mann, K. H., and J. R. N. Lazier. 1991. Dynamics of marine ecosystems:
biological-physical interactions in the oceans. Blackwell Scientific,
Boston.
16 Benthic Suspension Feeders and Flow
17
18 Benthic Suspension Feeders and Flow
Flow simulation
It is required to simulateflowso as to perform experiments in which flow
speed (unidirectional flow) orflowspeed and periodicity offlowreversal
(oscillating flows) are controllable variables. The flume requirements for
physiological and behavioral experiments are often less rigorous than the
full dynamic similarity required for realistically simulating sediment-
water interface conditions within flumes.
Although for convenience we have separated methods of simulating
flow into those in the field and those in the laboratory, the distinction is
not always easy to make. This is because there is a gradual change from
one to the other, with the difference in the degree of control of the
important variables: velocity, salinity, and temperature.
Field flows
Flow simulators suitable for ecological studies in the field have been
designed to measure growth rates of small populations of bivalves, or to
determine ecosystem level fluxes of materials, e.g. seston, plant nutri-
ents, carbon, or dissolved oxygen. The former are required to reproduce
natural flows, so that growth rate is a function of independent variables
such as bulk velocity or bottom shear stress ([/*), unhindered by wall
effects or an undeveloped boundary layer. For ecosystem fluxes, the
channels require well-defined flows so that the sampling at inlet and
outlet gives a good estimate of the fluxes of the materials of interest. All
of these are flowthrough devices and use a natural seawater supply. Flow
velocities used are either natural tidal currents or simulated ones using a
submersible pump or head tank and gravity flow.
The Kirby-Smith growth apparatus consists of eight 1.5-m-long acrylic
tubes which receive seawater from a constant level head tank (Fig. 2.1).
In our experiments (Wildish and Kristmanson 1985), bivalves were sup-
ported in the mid-centre line by plastic mesh inserts which minimally
impededflow.Flow rates through the tubes could be regulated by chang-
ing the diameter of the outlet tube fitting. The bulk flow velocities are
Methods of Study 19
A -d
I
I
lU
1 metre
(2.1)
where
U is the bulk velocity in cm s"1
Q is the volumetric flow in cm3 s"1
A is the cross sectional tube area, cm2
Pipe flows are well known hydrodynamically (Schlichting 1979). Mean
velocity profiles and turbulence characteristics in empty pipes can be
related to the mean pipe velocity and diameter. However, the pipes used
in the Kirby-Smith apparatus are too short for a fully developed profile
and the presence of inserts and bivalves disturbs theflow.Thus, the bulk
velocity measures recorded may not estimate the velocity experienced
by an experimental subject very well and may be apparatus-specific in
determining their responses.
20 Benthic Suspension Feeders and Flow
1.5 m 5m
!:
i ; 64"
if r.m ; o
\i i
ii I ; 45.7
ii i ; cm
g
Figure 2.2 Multiple-channel flume of Wildish and Kristmanson (1988): upper,
plan view; lower, elevation view; a, inlet port; b, perforated plywood; c, brass
screens; d, perforated bulkhead; e, adjustable weir; f, acrylic plastic (Plexiglas)
windows; g, drain.
Figure 2.3 The benthic ecosystem tunnel of Dame et al. (1984): a, acrylic sheet;
b, neoprene skirt; c, reinforcement rod.
Figure 2.4 Diagram of the Sylt flume. The plastic sheeting can be removed
when the flume is not in use (Asmus and Asmus 1991).
(2.2)
LW
where
Q is the volumetric flow over the bed in m3 h l
L is the distance, in m, between a and b
W is the path length of sediment, in m, over which the canalized
water flows
Methods of Study 23
This equation can only be used if the flow is well mixed with respect to
dissolved oxygen concentrations at the inlet and outlet. Otherwise, the
inputs and outputs, R, must be determined by integration of the form
R = \\vcda (2.3)
tical analysis was included. A further restriction for field flumes is that
they can only operate intertidally where the tidal prism is less than a few
metres, whereas tunnels could operate subtidally at least to SCUBA
diving depths.
JL
Tl U "b
1 metre
Figure 2.5 Vogel-LaBarbera recirculating flume in side view is constructed of
plywood and wooden beam supports: a, viewing window; b, outlet; c, collimators;
d, propeller shaft (Vogel and LaBarbera 1978).
of taxa (Table 2.1). Vogel (1981) pointed out that round pipe, rather than
the original square type of the Vogel and LaBarbera flume, is preferable,
and it has been used by subsequent authors listed in Table 2.1. All of the
flumes shown in Table 2.1 are recirculating, meaning that afixedvolume
of seawater is propelled through a closed, or partially open, circular
arrangement of conduits. If filtration rates are being measured on the
basis of the temporal change in seston concentration, the necessity is to
limit the volume appropriately to the filtration power of the suspension
feeder subjects. A corollary is that this limits the exposure time available
for experimental purposes with suspension feeders because of a buildup
of excretory substances.
Vogel-LaBarbera flumes are limited to flows ^SOcm-s"1, because at
greater velocities, air bubbles begin to appear as a result of cavitation
effects at the propeller. Because the flow entrance and exit conditions of
the Vogel-LaBarbera type flume (Fig. 2.5) do not reduce the instabilities
introduced by forcing the flow around four corners, it causes wave for-
mation, which limits its usefulness at higher flows. If higher flows are
required, then two flumes, the Saunders and Hubbard and Denny
high speed, are available - the former useful to 100 and the latter to
500 cm s \ The large Denny high speed flume is expensive to build, while
26 Benthic Suspension Feeders and Flow
the Saunders and Hubbard (Fig. 2.6) is mid-way in expense between the
Denny and Vogel-LaBarbera flumes. The design improvements of the
Saunders and Hubbard and Denny flumes which allow higher flows free
of waves and aeration were borrowed from naval architects (Saunders
and Hubbard 1944). For both flumes, they include a throat section choke
and turning vanes at each of the corners.
An annular, open channel flume or raceway was used by Emig and
Becherini (1970) and Leversee (1976), and this design appears to have
similar maximum velocity characteristics to the Vogel and LaBarbera
flumes.
The constant head flow tube or water tunnel of Charters and
Methods of Study 27
Wave character
Volumetric Oscillation umax
Flume type capacity (L) period (s) (cm-s-1) Reference
Lofquist >1000 3-25 30-150 Lofquist (1977)
Lofquist >1000 3-30 30-150 Hunter (1989)
Vogel-LaB arber a 15 l->3600 Trager et al. (1990)
Lofquist >1000 3-30 30-150 Turner and Miller
(1991), Miller et al.
(1992)
1m
in air. A field version of the annular flume - the Sea Carousel - from
Amos et al. (1992) was designed to measure in situ sediment erodibility
of shallow inter- or subtidal locations. For this purpose, it is equipped
with optical backscatter sensors, a video camera, and an electromagnetic
flow meter.
The interaction of siphonal exhalant currents generated by bivalves
with the development of the concentration boundary layer was studied
by introducing rhodamine dye into the exhalant flow by O'Riordan et al.
(1993). The effect of different boundary layer velocities on the concen-
tration field was determined by using a non-intrusive snapshot of the
dye concentration field. The dye was illuminated with a thin sheet of
laser light, and a charged coupled device camera was used to capture
images of the dye concentration. As the fluorescent light intensity is
proportional to the dye concentration, the method indicates rhodamine
concentration.
Velocity measurement
In order to choose a suitable velocity measurement system, answers to
some key questions must be obtained. The questions include, Is the
measurement to be of ambient velocity in the field or in the laboratory
flume?, At what spatial and temporal scales is velocity to be determined?
and At what hierarchical level of biological organization are the velocity
measurements to apply?
If the hierarchical level of biological organization of interest is ecologi-
cal, then the standard techniques of physical oceanography, such as
electromagnetic current meters, are applicable. Such current meters pro-
duce a continuous temporal record of velocity, and often current direc-
tion, stored electronically and retrievable with a laptop computer with-
out moving the sensor (Table 2.4). The interested biologist, by either
luck or design, may also be able to use a predictive physical oceano-
graphic model to input mean and maximum tidal velocities (e.g. Wildish,
Peer, and Greenberg 1986).
A limitation of electromagnetic current meters, and the models they
give rise to, is that they usually do not measure orbital motions due to
wind-wave effects. Studies by Denny (1982) on wave-exposed rocky
shores were based on transducers to measure the wave force received by
models of suspension-feeding organisms, e.g. barnacles. The force trans-
ducer output was converted to a frequency modulated (FM) radio signal
Methods of Study 31
Figure 2.8 Maximum force recorder of Bell and Denny (1994): a, spring; b,
rubber sleeve indicator of maximum force; c, upper plug; d, monofilament line; e,
practice golf ball.
v 5 7
dt ~ v^# v
where
m = mass of the cast in kg at time t
K = m-s"1, rate constant for a specified shape
A = m2, area exposed to flow
Cs = kg m3, saturation constant of gypsum
C = kg m3, concentration of dissolved gypsum
Methods of Study 33
j ( ) (2.5)
where
F = transfer rate of gypsum per unit area, kg m2 s"1
D = diffusion coefficient for CaSO4 in water, m2 s"1
8 = concentration boundary layer thickness, m
The dissolution rate of gypsum is sensitive to temperature for two rea-
sons: The diffusivity increases with temperature and decreases in vis-
cosity are caused by increases in temperature, which, in turn, reduce
boundary layer thickness, 8.
A number of shapes and materials have been used to measure dissolu-
tion rates - small rings, actually candy (Lifesavers) - sewn to a kelp
surface (Koehl and Alberte 1988) to "clod cards" of gypsum cast in
plastic ice cube trays (Doty 1971; Jokiel and Morrissey 1993; Thompson
and Glenn 1994) to cylinders (Wildish et al. 1997). The ideal shape for
such measurements is a sphere, since flow from any direction should
equally influence the dissolution process. A capped right cylinder will
also be insensitive to the direction of flow as long as its main axis is
oriented perpendicular to the flow. The weight initially (Mo) and after a
measured exposure period, drying, and reweighing (Mx) may be used
in a comparative or absolute measure of water movement. The
dimensionless index proposed by Doty (1971), the diffusion factor, DF,
is calculated as
[ J flow
(2.6)
still
where the weight loss, determined in the field for a standard time, is
divided by the weight loss in still seawater of the same salinity and
temperature.
In order to relate gypsum cylinder weight losses to flow in an absolute
measure of velocity, it is necessary to calibrate flume flows and relate
weight loss to salinity, temperature, and velocity. In the field, replicate
cylinders may be deployed (Fig. 2.9) on a frame which includes a swivel
and vane so that the gypsum cylinders face the current.
Five general techniques have been used to measure velocity in labora-
34 Benthic Suspension Feeders and Flow
used to confirm that velocity at the test point was equal to the volumetric
flow divided by the orifice area. Calibration plots are made of probe
output in hertz as a function of volumetric flow rate.
A similar but miniaturized version of the design was used by
Muschenheim et al. (1986) to calibrate small thermistor bead probes
(Fig. 2.11).
Seston sampling
The purpose of seston sampling is to determine the concentration and
composition of seston particles in seawater. During feeding by suspen-
sion feeders, it is necessary to take grab samples of seawater with small
diameter sampling tubes, because of the localized nature of the concen-
tration gradient of seston within the benthic boundary layer.
Three general types of seston sampling may be used to determine
seston concentration and composition: isokinetic, non-isokinetic, and
active suspension feeder-biased. Non-isokinetic is the default option,
where sampling is arbitrary, not considering flow and concentration
boundary layer conditions. Isokinetic is the appropriate means of sam-
pling for passive suspension feeders, and it, or species-biased sampling to
simulate the live animal, is appropriate for suspension feeders with an
active ciliary or muscular pump.
Isokinetic sampling
A theory of sampling solid particles from flowing fluids has been devel-
oped for aerosols in industrial applications (Hinds 1982). Sampling er-
Methods of Study 39
rors may arise in sampling flowing fluids to measure the solids content as
a result of the following:
A mismatch between ambient and sampling tube velocities
Variations in settling velocities of the sestonic particles
The orientation of the sampling tube with respect to the flow
With regard to the latter, Hinds (1982) showed that as the orientation of
the sampling tube was moved from opposing the flow (0) to an increas-
ing angle until it was normal to the flow (90), so sampling errors due to
mismatch of flows also increased. Estimates of this error are given in
graphical form in Hinds (1982).
The key hydrodynamic parameter for sampling is the dimensionless
Stokes number, Stk
W JJ
Stk = ^ ^ (2.7)
where
Ws = passive terminal settling velocity of seston, cm s"1
Uo = ambient velocity, cm s"1
g = gravitational constant, 981 cm s~2
D = diameter of sampling tube, cm
The Stokes number indicates the likelihood that the particle will follow
the fluid streamlines, with higher values of Stk giving greater sampling
error. Thus, if Stk < 0.01 and the velocity mismatch between ambient
and sampling flow is within 4 times the isokinetic rate, then the sam-
pling error will be <5% (Hinds 1982). The maximum terminal settling
velocity of a sestonic particle at a fixed Stk = 0.01 can be calculated
from
W,<9-f (2.8)
Assuming D = 0.64 cm, we can calculate at Uo = Scm-s"1 that seston with
a settling velocity of -l^Scm-s" 1 will be collected without significant
sampling error. Similarly, at lOcm-s"1, but otherwise the same condi-
tions, the limit for collection bias is particles of settling velocity of
-0.63 cm-s"1. These considerations suggest that sestonic particles such as
typical phytoplankton cells with Ws < 0.01 cm s"1 would be collected
without sampling error (see also Chap. 5). However, if the seston load
contained denser sedimentary particles >100|mn, then sampling errors, if
strict isokinetic sampling were not practised, might occur.
40 Benthic Suspension Feeders and Flow
Table 2.5. Matching flume flows, Uo, and sample tube flow at the
iso kinetic poinC
a
Tubing is D = 0.140 cm, A = 0.0154 cm2. PE200 Clay-Adams Intramedic
tubing, Becton and Dickson, Parsipanny, NJ 07054, USA. Static head required
to deliver matched flows is shown with the time required to empty the total
tube volume completely.
Non-isokinetic sampling
From the preceding discussion, we conclude that non-isokinetic sampling
will rarely lead to serious sampling errors in collecting seston. Only
where sedimentary particles are resuspended during feeding obser-
vations or experiments, as was the case in a study of a facultative
Methods of Study 41
Seston collection
Before a quantitative estimate of seston concentration can be made, it is
necessary to filter the sample - all or a suitable subsample - so that the
result can be given as dry weight per unit volume of seawater. Other
measures of concentration commonly used include chlorophyll a, deter-
mined either fluorometrically or spectrophotometrically (Strickland and
Parsons 1968); adenosine triphosphate (ATP), as an indication of all
living cells or tissue (Holm-Hansen and Booth 1966); and the acridine
orange epifluorescence direct counting technique of Hobbie, Daley,
and Jaspar (1977) to indicate numbers of live bacteria (Wildish
and Kristmanson 1984). Electronically sizing and counting sestonic par-
ticles (Strickland and Parsons 1968) are other frequently used options
in determining seston concentration. More recently, the wider availa-
bility of the flow cytometer has meant that this technique has been
used (e.g. Shumway et al. 1985; Boucher, Vaulot, and Partensky 1991)
to measure particle numbers and fluorescence simultaneously. Shumway
et al. (1985) showed that even in mixed phytoplankton cultures con-
taining three species, this technique could indicate the proportional
presence of each species. The flow cytometric technique as it applies
to seston in seawater is reviewed by Yentsch et al. (1983); also see
the review of particle measurement in seawater edited by Demers
(1991).
All of the measures of seston concentration described - except the
simplest, dry weight determination - are selective, and none of them is
capable of indicating the quality of the seston as a food source to a
named suspension feeder.
Methods of Study 43
Seston quality
In a study of blue mussel initial feeding responses with a unialgal culture,
Wildish and Miyares (1990) found that two factors influenced filtration
rate - velocity and chlorophyll a content per cell. In these experiments,
the mussels were all fed the same initial cell density of the cultured
microalga, but, as a result of growth/time related differences in different
cultures, the chlorophyll a content varied, with evidence that the mussels
preferred those cells with higher levels of chlorophyll a. This same
method cannot be used with a mixed culture or with wild-caught
phytoplankton because of differences in amounts of chlorophyll a typical
of a species as well as the intraspecific variances already noted. Use of
flow cytometric techniques may allow more species of cultured
microalgae in a mixed culture to be resolved (e.g., see Olson and Zettler
1995), even where preferential clearance of one species over another is
present. However, it too may be of limited use in field samples, where
many phytoplankton species are usually present and only a few of them
are being ingested.
The nutritional quality of seston for suspension feeders could not be
adequately expressed by gross biochemical measures of the seston
(Navarro and Thompson 1995). Navarro and Thompson proposed a
dimensionless food index based on the proportion of protein, carbohy-
drate, and lipid content of the total seston available. They demonstrated
seasonal changes of the food index, but did not relate them to the growth
of any named suspension feeders.
The nutritional value of seston to suspension feeders can only be
determined in standardized feeding growth bioassays. The availability of
methods to create microparticulate diets (Jones, Mumford, and Gabbott
1974; Langdon, Levine, and Jones 1985; Southgate, Lee, and Nell 1992)
with known types and quantities of nutrients suggests that for larvae,
juveniles, and adults it is possible to create a standard diet for genetically
standardized suspension feeders. Such a diet would be of great value in
comparing the quality of the available natural seston as a food for sus-
pension feeders.
Specialized methods
Many different methods have been used in studying the larval biology,
physiology, ethology, and ecology of benthic suspension feeders. Be-
44 Benthic Suspension Feeders and Flow
cause there are so many, we have selected a few methods for each of the
four disciplines concerned and present them here.
1 0
0
o
o
o
b
c
(
71
1 cm
Figure 2.13 Miniature flow tank for observing tethered larvae (after Emlet
1990): a, air lift; b, upstream reservoir; c, downstream reservoir; d, inlet with
control valve; e, collimator; f, larva; g, coverslip; h, tether; i, exit port with control
valve.
R = P(RE) (2.9)
Seston particles which differ in size or chemical composition will vary in
retention efficiency in bivalves (Famme and Kofoed 1983), so that the
clearance rate can be said to be particle-type-specific.
Historically, indirect methods for measuring S or R have been most
commonly used. Such measurements have been made in closed tanks
of fixed volume ("static") or in variously shaped flowthrough vessels.
The difference between the initial, Q, and final, C2, concentrations
of seston, or the inlet and outlet concentrations is used to estimate S
or R. The volume of the static system, V, and the mass of animals, M,
must be known as well as the elapsed time of the experiment, T, to
calculate S:
(2.10)
MT
Typical units for this equation might be g-seston-g"1 dry weight -h"1.
The clearance rate for a unit weight of animal can be estimated from
R = SC~1 {111)
with reasonable accuracy if Cx and C2 are not greatly different (say,
C2 > 0.9 C\). Unacceptable inaccuracies may arise because the equation
is based on the assumption that the seston concentration of the filtered
seawater remains at Cx during the experiment. Hildreth and Crisp (1976)
recognized this difficulty and suggested that Q in equation (2.11) be
replaced by the actual seston concentration around the suspension
feeder throughout the experimental test.
If Q divided by C2 is relatively large, the formula of Coughlan (1969)
is preferable:
Figure 2.14 Direct method of determining the pumping rate of a mussel (after
Famme et al. 1986): a, laser source; b, mirror; q and c2, inhalant and exhalant
sides of a divided chamber; d, membrane; e, shunt; f, pump; g, scale.
Jones and Allen (1986) describe a pressure sensor with a <l-mm tip for
sampling inhalant/exhalant pressures. Because of the complex flow and
pressurefieldswithin the exhalantflow,it is critical to position the sensor
precisely. Because of these exacting requirements, such methods have
not often been successfully used.
Mohlenberg and Riisgard (1978) describe an apparatus which can be
used in non-flowing conditions to sample the exhalant flow of active
suspension feeders isokinetically. The sampling flow is adjusted so that
the seston concentration is minimal, at which point the flows are consid-
ered to be isokinetic. Mohlenberg and Riisgard's (1978) method could
not be used satisfactorily in flume experiments because of the complicat-
ing ambient flow around the bivalve (Wildish and Saulnier 1993); direct
exhalant flow sampling (Wildish and Saulnier 1993) was substituted. The
bivalve gill retention efficiency (RE) can be directly determined from the
concentration of inhalant (Q) compared to exhalant seston (C2):
Methods of Study 51
RE = 1 - ^ - (2.13)
c1
The value of R can be calculated as in equation (2.11) and the pumping
rate, P, can be calculated by dividing R by RE, provided the estimates
are for the same periods.
Suspension feeding bioassays ideally require that seston be main-
tained at a constant concentration during the experiment. Winter (1973)
originally used a gently stirred tank of 10-L capacity containing blue
mussels. Improvements to this technique have been suggested by
Riisgard and Mohlenberg (1979) and a less expensive version proposed
by Haupt (1979). All of these methods use a photometer to monitor the
particle concentration by differences in light transmission, without dis-
tinguishing the preferred microalga from other sestonic particles offered
to the mussel. As the mussels consume the seston, the increasing light,
through a relay, switches on a dosing valve connected to a microalgal
culture for a predetermined period. This allows the culture to be added
to the tank at a rate to maintain a pre-set seston concentration. The rate
of addition of the microalgae is also a direct measure of mussel uptake
or clearance, after a suitable allowance for microalgal settling or growth
and cell division has been made in a control run with mussels absent
but for the same period as the experiment. Photometric determination
is not specific for microalgal particles and any dense particle, such
as faeces or pseudofaeces, reduces the light reaching the photoelectric
sensor.
A more specific method for microalgal cells involves fluorescence
measurements with two suitable filters, providing exciting light at 430 nm
and emitted light at 650 nm, which is suitable for continuous measure-
ment of in vivo chlorophyll concentration (Lorenzen 1966) in a continu-
ous flow fluorometer (Table 2.6; Fig. 2.15). Some limitations of the in
vivo fluorescence method should be kept in mind (Loftus and Seliger
1975), e.g. high concentration quenching or direct absorption of light by
the microalgae, when using this method. Bayne, Widdows, and Newell
(1977) were successful in using flow fluorometry to measure the feeding
rate of blue mussels fed unialgal cultures. A suitable setup for maintain-
ing a constant seston concentration with feeding bivalves present in a
flume flow is shown in Fig. 2.15. One unsatisfactory feature of this
method is that, because of the different fluorescent properties of each
microalgal species, it is not possible to maintain the strict relationship
52 Benthic Suspension Feeders and Flow
Type Supplier
Temperature measurement Hewlett Packard, 6877 Goreway Drive,
Ontario, Canada L4V 1M8
Photoperiod control Suntracker, Paragon Electric Canada Ltd., 221
Evans Ave., Etobicoke, Ontario, Canada M8Z
1J5
Three-dimensional Velmex Inc., East Bloomfield, NY, 14443 USA
positioner made from
Unislide assembly parts
Spaghetti tags for identifying Floy Tag & Manufacturing Inc., 4616 Union
experimental subjects, Bay Place, N.E., Seattle, Washington, 98105
gypsum cylinders, etc. USA
Flow fluorometer, AU-005 Turner Designs, 845 W. Maude Ave.,
Sunnyvale, California, 94086 USA
Melamime formaldehyde Radiant Color, Richmond, California, 94809
particles USA
Endoscope Olympus Corp., Medical Instrument Div., 4-T
Nevada Drive, Lake Success, New York, 11042
USA
Sony video system with PEMTEK Technologies, 201 Brownlow 33B,
miniature camera Dartmouth, Nova Scotia, Canada B3B 1W2
Aerolab, 9580 Washington Blvd., Laurel, MD,
Schlieren systems 20723 USA
Yellow Springs Instrument Co. Inc., Box 279,
Dissolved oxygen probes Yellow Springs, Ohio, 45387 USA
b
d
Conversions
Wet weight, g, to kcal, multiply by:
1.0 (Steele 1974)
0.6 x fresh weight (Mills and Fournier 1979)
0.5 (Crisp 1975)
kcal to kJ, multiply by:
4.186
kJ to g carbon, divide by:
50
Predictions
Production/biomass, P:B, ratios may be determined by multiplying W, wet
body mass at maturity, g, or lifespan in years, based on the following
equations:
Applicability Equation Reference
Macroafauna + meiofauna P:B = 0.971 W^167 Schwinghamer (1981)
Macrofauna P :B = 0.563 W'0302 Schwinghamer (1981)
Invertebrates, 5-20C P:B= 0.640W0310 Banse and Mosher (1980)
Invertebrates, less insects P:B = 0.6UW0390 Banse and Mosher (1980)
Macrofauna P:B = 4.571 lifespan Robertson (1979)
References
Amos, C. L., J. Grand, G. R. Daborn, and K. Black. 1992. Sea carousel: a
benthic annular flume. Estuar. Coast. Shelf Sci. 34: 557-578.
Amouroux, J. M., M. Revault d'Allonnes, and C. Rouant. 1975. Sur la mesure
directe du debit de filtration chez les mollusques lamellibranches. Vie
Milieu (Ser. 2B) 25: 339-346.
Asmus, H., R. M. Asmus, T. C. Prins, N. Dankers, G. Frances, B. Maas, and
K. Reise. 1992. Benthic-pelagic flux rates on mussel beds: tunnel and tidal
flume methodology compared. Helgolander Meeresunters. 46: 341-361.
Asmus, R. M., and H. Asmus. 1991. Mussel beds: limiting or promoting
phytoplankton? J. Exp. Mar. Biol. Ecol. 148: 215-232.
Banse, K., and S. Mosher. 1980. Adult body mass and annual production/
biomass relationships of field populations. Ecol. Monogr. 50: 355-379.
Bayne, B. L., J. Widdows, and R. I. E. Newell. 1977. Physiological
measurements on estuarine bivalve molluscs in the field, p. 57-68. In B.F.
Keegan, P. 6. Ceidigh, and P. J. S. Boaden (ed.) Biology of benthic
organisms. Pergamon, Oxford.
Bell, E. C, and M. W. Denny. 1994. Quantifying "wave exposure": a simple
device for recording maximum velocity and results of its use at several
field sites. J. Exp. Mar. Biol. Ecol. 181: 9-29.
Best, B. A. 1988. Passive suspension feeding in a sea pen: effects of ambient
flow on volume flow rate and filtering efficiency. Biol. Bull. 175: 332-342.
Boucher, N., D. Vaulot, and F. Partensky. 1991. Flow cytometric
determination of phytoplankton DNA in cultures and oceanic
populations. Mar. Ecol. Prog. Ser. 71: 75-84.
Bowman, G. T., and J. J. Delfino. 1980. Sediment oxygen demand techniques:
a review and comparison of laboratory and in situ systems. Water Res. 14:
491^99.
Boynton, W. R., W. M. Kemp, C. G. Osborne, K. R. Kaumeyer, and M. C.
Jenkins. 1981. Influence of water circulation rate on in situ measurements
of benthic community respiration. Mar. Biol. 65: 185-190.
Bradshaw, P. 1964. Experimental fluid mechanics. Pergamon Press, Oxford.
Buchanan, J. B., and R. M. Warwick. 1974. An estimate of benthic
macrofaunal production in the offshore mud of the Northumberland
coast. J. Mar. Biol. Assoc. U.K. 54: 197-222.
Butman, C. A. 1986. Sediment trap biases in turbulent flows: results from a
laboratory flume study. J. Mar. Res. 44: 645-693.
Butman, C. A., J. P. Grassle, and C. M. Webb. 1988. Substrate choices made
by marine larvae settling in still water and in aflumeflow.Nature 333:
771-773.
Carey, D. A. 1983. Particle resuspension in the benthic boundary layer by a
tube-building polychaete. Can. J. Fish. Aquat. Sci. 40: 301-308.
1989. Fluorometric detection of tracer particles used to study animal-
particle dynamics. Limnol. Oceanogr. 34: 630-634.
Charriaud, E. 1982. Direct measurements of velocity profiles and fluxes at the
cloacal siphon of the ascidian Ascidiella aspersa. Mar. Biol. 70: 35-40.
Methods of Study 61
cavity in Mytilus edulis (L.) and its influence on particle retention. Mar.
Biol. Lett. 4: 207-218.
Famme, P., R. U. Riisgard, and C. B. J0rgensen. 1986. On direct
measurements of pumping rates in the mussel, Mytilus edulis. Mar. Biol.
92: 323-327.
Foster-Smith, R. L. 1976. Pressures generated by the pumping mechanism of
some ciliary filter feeders. J. Exp. Mar. Biol. Ecol. 25: 199-206.
Gallagher, S. M. 1988. Visual observations of particle manipulation during
feeding in larvae of a bivalve mollusc. Bull. Mar. Sci. 43: 344-365.
Galstoff, P. S. 1928. Experimental study of the function of the oyster gills and
its bearing on the problems of oyster culture and sanitary control of the
oyster industry. Bull. U.S. Bur. Fish. 44, Doc. 1035: 1-39.
Grassle, J. P., P. V. R. Snelgrove, and C. A. Butman. 1992. Larval habitat
choice in still water and flume flows by the opportunistic bivalve Mulinia
lateralis. Neth. J. Sea Res. 30: 33-44.
Grizzle, R. E., R. Langan, and W. H. Howell. 1992. Growth responses of
suspension-feeding bivalve molluscs to changes in water flow: differences
between siphonate and nonsiphonate taxa. J. Exp. Mar. Biol. Ecol. 162:
213-228.
Gruffyd, L. D. 1976. Swimming in Chlamys islandica in relation to current
speed and an investigation of hydrodynamic lift in this and other scallops.
Nor. J. Zool. 24: 365-378.
Gust, G. 1982. Tools for oceanic small-scale high-frequency flows: metal clad
hot wires. J. Geophys. Res. 87: 445-447.
Hannan, C. A. 1984. Planktonic larvae may act like passive particles in
turbulent near-bottom flows. Limnol. Oceanogr. 29: 1108-1115.
Hargrave, B. T., and W. M. Burns. 1979. Assessment of sediment trap
collection efficiency. Limnol. Oceanogr. 24: 1124-1136.
Hargrave, B., G. Siddal, G. Steeves, and G. Await. 1994. A current-activated
sediment trap. Limnol. Oceanogr. 39: 383-390.
Hartnoll, R. G. 1967. An investigation of the movement of the scallop. Helg.
Wiss. Meeresunters. 15: 523-533.
Haupt, K. 1979. A simple device to control concentrations of food algae in
feeding experiments with filter feeding organisms. Veroff. Inst.
Meeresforsch, Bremerhaven 17: 241-244.
Hildreth, D. I., and D. J. Crisp. 1976. A corrected formula for calculation of
filtration rate of bivalve molluscs in an experimental flowing system. J.
Mar. Biol. Assoc. U.K. 56: 111-120.
Hinds, W. C. 1982. Aerosol technology. John Wiley & Sons, New York.
Hobbie, J. E., R. Daley, and S. Jasper. 1977. Use of nucleopore filters for
counting bacteria by fluorescence microscopy. Appl. Environ. Microbiol.
33: 1225-1228.
Holm-Hansen, O., and C. R. Booth. 1966. The measurement of adenosine
triphosphate in the ocean and its ecological significance. Limnol.
Oceanogr. 11: 510-519.
Huettel, M., and G. Gust. 1992. Impact of bioroughness on interfacial solute
exchange in permeable sediments. Mar. Ecol. Prog. Ser. 89: 253-267.
Hunt, M. J., and C. G. Alexander. 1991. Feeding mechanisms in the barnacle
Tetraclita squamosa (Brugiere). J. Exp. Mar. Biol. Ecol. 154: 1-28.
Hunter, T. 1989. Suspension feeding in oscillating flow: the effect of colony
morphology and flow regime on plankton capture by the hydroid Obelia
longissima. Biol. Bull. 176: 41-49.
Methods of Study 63
Vogel, S., and N. Feder. 1966. Visualization of low-speed flow using suspended
particles. Nature 209: 186-187.
Vogel, S., and M. LaBarbera. 1978. Simple flow tanks for research and
teaching. Bioscience 28: 638-643.
Wainwright, S. C. 1990. Sediment-to-water fluxes of particulate material and
microbes by resuspension and their contribution to the planktonic food
web. Mar. Ecol. Prog. Ser. 62: 271-281.
Walne, P. R. 1972. The influence of current speed, body size and temperature
on the filtration rate of five species of bivalves. J. Mar. Biol. Assoc. U.K.
52: 345-374.
Ward, J. E., B. A. MacDonald, and R. J. Thompson. 1993. Mechanisms of
suspension feeding in bivalves: resolution of current controversies by
means of endoscopy. Limnol. Oceanogr. 38: 265-272.
Wildish, D. J., and D. D. Kristmanson. 1979. Tidal energy and sublittoral
macrobenthic animals in estuaries. J. Fish. Res. Board Can. 36: 1197-
1206.
1984. Importance to mussels of the benthic boundary layer. Can. J. Fish.
Aquat. Sci. 41: 1618-1625.
1985. Control of suspension-feeding bivalve production by current speed.
Helgol. Wiss. Meeresunters. 39: 237-243.
1988. Growth response of giant scallops to periodicity of flow. Mar. Ecol.
Prog. Ser. 42: 163-169.
Wildish, D. J., and U. Lobsiger. 1987. Three dimensional photography of soft
sediment benthos, S.W. Bay of Fundy. Biol. Oceanogr. 4: 227-241.
Wildish, D. J., and M. P. Miyares. 1990. Filtration rate of blue mussels as a
function of flow velocity: preliminary experiments. J. Exp. Mar. Biol.
Ecol. 142: 213-219.
Wildish, D. J., and D. Peer. 1981. Method for estimating secondary production
in marine Amphipoda. Can. J. Fish. Aquat. Sci. 38: 1019-1026.
Wildish, D. J., D. L. Peer, and D. A. Greenberg. 1986. Benthic macrofaunal
production in the Bay of Fundy and the possible effects of a tidal power
barrage at Economy Point-Cape Tenny. Can. J. Fish. Aquat. Sci. 43:
2410-2417.
Wildish, D. J., and A. M. Saulnier. 1993. Hydrodynamic control of filtration in
Placopecten magellanicus. J. Exp. Mar. Biol. Ecol. 174: 65-82.
Wildish, D. J., J. D. Trynor, and H. M. Akagi. 1997. Measurement of water
movement by gypsum dissolution method in the Bay of Fundy. Can. Tech.
Rep. Fish. Aquat. Sci. (in press).
Winter, J. E. 1973. The filtration rate of Mytilus edulis and its dependence on
algal concentration, measured by a continuous automatic recording
apparatus. Mar. Biol. 22: 317-328.
Wolaver, T., G. Whiting, B. Kjerfve, J. Spurrier, H. McKellar, R. Dame, T.
Chrzanowski, R. Zinmark, and T. Williams. 1985. The flume design: a
methodology for evaluating material fluxes between a vegetated salt
marsh and the adjacent tidal creek. J. Exp. Mar. Biol. Ecol. 91: 281-291.
Wright, R. T., R. B. Coffin, C. P. Ersing, and D. Pearson. 1982. Field and
laboratory measurements of bivalve filtration of natural marine
bacterioplankton. Limnol. Oceanogr. 27: 91-98.
Yentsch, C. M., P. K. Horan, K. Muirhead, Q. Dortch, E. Haugen, L.
Legendre, L. S. Murphy, M. J. Perry, D. A. Phinney, S. A. Pomponi, R.
W. Spinrad, M. Wood, C. S. Yentsch, and B. J. Zahuranic. 1983. Flow
68 Benthic Suspension Feeders and Flow
cytometry and cell sorting: a technique for analysis and sorting of aquatic
particles. Limnol. Oceanogr. 28: 1275-1280.
Young, R. A. 1977. Seaflume: a device for in situ studies of threshold erosion
velocity and erosional behaviour of undisturbed marine muds. Mar. Geol.
23: M11-M18.
Yund, P. O., S. D. Gaines, and M. D. Bertness. 1991. Cylindrical tube traps for
larval sampling. Limnol. Oceanogr. 36: 1167-1177.
3
Dispersal and settlement
Study of larval biology has been under way for -150 years and a large
literature is available, causing Young (1990) to suggest that ideas are
often rediscovered by each new generation of scientists interested in this
field. This literature has been repeatedly reviewed; the most useful and
recent include Crisp (1974), Chia, Buckland-Nicks, and Young (1984),
Scheltema (1986), Butman (1987), Pawlik (1992), Sammarco and Heron
(1994), and McEdwards (1995). The overview presented here relies
heavily on these publications and on selected references which have
appeared in the decade subsequent to 1985. The basis for selection of
references in this chapter is that only those with flow-related content are
included.
Consistent with the aims outlined in Chapter 1, we emphasize hydro-
dynamic mechanisms in larval biology where these are known. Conse-
quently, our concern here is limited to larval and post-larval dispersal at
spatial scales >0.1km and on the environmental factors, notably hydro-
dynamic ones, which influence larval settlement/recruitment at spatial
scales <0.1km on both soft and hard substrates. We begin with back-
ground coverage of the typical benthic suspension feeder life cycle and
dispersal in general.
69
70 Benthic Suspension Feeders and Flow
Adult
Juvenile
Figure 3.1 Life cycle diagram of the acorn barnacle, Semibalanus balanoides
(based on Pyefinch 1948).
produce the post-larval stage or juvenile and, eventually, the adult barna-
cle. Barnacles are hermaphroditic and cross fertilization, by means of an
intromittent organ, involves internal fertilization, followed by release of
the first larval stage to begin the pelagic phase of life. There is a great
range in shapes and sizes of larvae among suspension feeding inverte-
brates, although not all, e.g. polychaete larvae, have recognizably distinct
stages as, for example, do barnacles. Some polychaetes simply add seg-
ments to the body as larval life in the plankton proceeds.
The duration of pelagic life for larval suspension feeders is dependent
on both genetic factors and environmental variables. A characterization
Dispersal and Settlement 71
based on whether they feed during planktonic life may indicate the
approximate duration of larval life (Table 3.1). Thus, lecithotrophic lar-
vae are short-lived (0.1-2 d) and do not feed, whereas planktotrophic
larvae have the potential for a longer life (e.g. some teleplanic larvae ~2
years) and may feed at various periods during the larval stage (Sheltema
1986). The length of larval life in the plankton generally increases from
type 1 through type 4 in Table 3.1.
Competent larvae are those with the ability to settle on a soft or hard
substrate in the general way shown in Fig. 3.2. According to this scheme,
the competent larva may make several contacts with the substrate, by
either passive sinking or active swimming, before temporarily or perma-
nently burrowing or attaching to it. Whether attachment or burrowing is
involved after initial contact depends on the competent larva receiving
ecologically relevant cues at sensory receptors. If appropriate environ-
mental cues are present, the larvae undergo a hormonally coordinated
physiological change, or metamorphosis, to produce the post-larval stage,
which may precede, coincide with, or follow settlement (Scheltema 1974;
Butman 1987). Settlement simply refers to the change from planktonic to
benthic life of the competent larva; it may involve repeated explorations
72 Benthic Suspension Feeders and Flow
Competent larva
Temporary attachment
crawling v
activity >v -
\ nervous sensory Environmental
/ system ~ "receptor" cue
metamorphic factor
unmasked
Post-larva
Dispersal
Dispersal is the general process by which organisms become distributed
spatially, at scales generally >0.1km, to potentially hospitable niches.
Scheltema (1986) outlined four main types of dispersal common among
benthic invertebrates:
Most of this chapter deals with larval and post-larval dispersal, although
we first briefly consider rafting and synanthropic dispersal. The purpose
of this is to ensure for the reader that these forms of dispersal are not
forgotten when pertinent questions regarding it are considered and dis-
persal hypotheses framed.
Dispersal by rafting occurs on a wide variety of living and dead floating
substrates, which may carry all life history stages of suspension feeders
wherever the surface currents dictate. Descriptions of the associated
fauna with surface drifting macroalgal rafts stress the importance of a
resident community, e.g. off the coast of Florida (Stoner and Livingston
1980). DeVantier (1992) described rafting of tropical marine animals,
inclusive of reef and pearl oysters, as well as species of Bryozoa, on the
buoyant skeleton of the reef coral Symphylla agarica, in which chambers
in the coral were air-filled, giving lift during rafting. Other species of
macroalgae may be submerged and moved by benthic currents along the
seabed, rather like marine tumbleweeds. Holmquist (1994) reported that
Laurencia poiteau does this and can therefore carry rafted fauna to new
benthic niches.
The importance of synanthropic dispersal of a wide range of marine
74 Benthic Suspension Feeders and Flow
Larval dispersal
Planktotrophic larvae are generally <1000|im in length and swim/sink at
Reynolds numbers in the creeping flow range. Such larvae swim by
means of beating cilia, although larvae which are longer than 1000 Jim
are unable to swim in this way as a result of inefficiencies linked to size,
and are forced to use muscular contractions to swim (Chia et al. 1984).
Examples include the later stages of some polychaete larvae or the
lecithotrophic tadpole larvae of ascidians. Chia et al. (1984) showed that
larvae using muscular power to swim increased their speed proportional
to an increase in body size, although this was not the general case with
ciliary swimming larvae.
Although data are readily available for larval size in the literature (for
example, see the review of larval stages of bivalves presented by
Ackerman et al. 1994), concomitant and reliable data on swimming
Dispersal and Settlement 75
speed and sinking rate are often lacking, explaining the few entries in
Table 3.2.
The key factors which influence the final distances achieved by larvae
during dispersal and before settlement of the competent stage are dura-
tion in the plankton, larval sinking rates, larval swimming speed, and
direction and velocity of ambient currents. Environmental factors which
influence the length of larval life, and therefore the potential dispersal
distance, are physical, e.g. salinity and temperature, or biological, e.g. the
availability and quality of sestonic food or predation.
The length of time that larvae are planktonic, both between species
and within species, is highly variable. Hence, lecithotrophic larvae may
spend a few hours or days in the plankton, with the outcome that they
either settle or die because of predation or limiting energy resources
(Scheltema 1986). Longer-lived planktotrophic larvae are affected by
physical factors in various ways. As an example, the dispersal of the
larval stages of the blue mussel, following the release of sperm and egg,
is depicted in Fig. 3.3. After fertilization, the trocophore is followed by
the veliger, the velichoncha, the eyed veliger, and the pediveliger, which
is the competent settling stage, and this sequence can take from 26 to 60 d
(Bayne 1964a). After settling on seaweed, the pediveliger metamor-
phoses to the first post-larval or spat stage, termed the plantigrade (not
shown in Fig. 3.3). The plantigrade can actively migrate to a more suit-
able hard substrate nearby to grow into the juvenile and adult forms of
the blue mussel. Bayne (1965) conducted growth experiments with blue
mussel pediveligers in which optimum salinities of 30-33 parts per thou-
sand (%o) and food concentrations of 1 x 105 cells L"1 were offered while
varying the temperature. It was found that mixed cultures of microalgae
led to better growth rates than if a unialgal culture was offered as food
and that the delay of metamorphosis achieved by pediveligers ranged
from ~2 to 46 d, in inverse relation to temperature (Bayne 1965). Salinity
also affected the time spent as a pediveliger, so that the delay was greater
at the lower salinities tested of 12.5%o (Bayne 1965).
From a knowledge of the duration of planktonic life it is possible to
predict crudely the distances the larvae could disperse on the basis of the
dispersion diagrams prepared by Okubo (1971). A three-dimensional
mathematical model of ocean circulation over Georges Bank was con-
structed by Tremblay and Sinclair (1994) to predict the passive transport
of larvae of the giant scallop Placopecten magellanicus. The model was
reasonably successful in predicting the flow field and, consequently, in
simulating the scallop larval drift over Georges Bank. Important vari-
Table 3.2. Larval size, swimming speed, sinking rates, and duration of stay in the plankton for some benthic
suspension feeders.
Size, urn
Sinking Swimming speed, cm s
Competent rate, Duration,
Taxon Egg larva cm s"1 Up Down Horizontal days Reference
Cnidaria, Octocorallia
Eunicella stricta 0.38 0.2 Theodor (1967)
Annelida, Polychaeta
Phragmatopoma 0.07-0.16 0.20-0.50 Pawlik et al.
lapidosa californica (1991)
Streblospio benedicti 300-1400 0.01-0.30 Butman (1989)
Mollusca, Bivalvia
Crassostrea virginica 240-320 -0.83 0.13-0.27 0.13-0.27 0.01-0.1 Hidu and Haskins
(1978)
Pecten maximus 66-70 240-260 0.10 0.12 32-41 Cragg (1980)
C. virginica >250 0.037-0.102 >12-24 Mann (1988)
Spisula solidissima 196.1 0.22 0.03-0.04 24 Mann et al. (1991)
Mulina lateralis 159.7 0.13 0.03 0.02 19 Mann et al. (1991)
Rangia cuneata 168.5 0.17 0.02-0.05 0.06 12 Mann et al. (1991)
Crustacea, Cirripedia
Balanus crenatus 0.17-0.55 12-20 DeWolf (1973)
Dispersal and Settlement 11
L3
straight hinge veliger
L2
young veliger L4
veliconcha
L1
trochophore
fertilized L5
zygote eyed
veliger
Figure 3.3 Diagram of larval dispersal and settlement of the blue mussel
Mytilus edulis (based on photomicrographs andfiguresin Bayne 1976). The post-
larval plantigrade is not shown.
10-
E
o 5-
Q
LU
LU
CL
CO 2-
O
1-
CO
0.5 -
60 100 140 180 220 260 300
MEAN LENGTH, pm
Figure 3.4 Effect of larval length of Crassostrea virginica on the log of swim-
ming speed at S = 25 o/oo and three different temperatures (Hidu and Haskins
1978). To obtain the size range shown, trocophores, straight hinge, and eyed
veligers were used. (From Hidu and Haskins 1978, reused by permission, Copy-
right Estuarine Research Federation.)
depth to determine larval swimming speeds of scallop. Butman (1989)
and Pawlik, Butman, and Starczak (1991) measured sinking rates with
narcotized or dead larvae and swimming speeds on live, suspension-
feeding polychaete larvae in settling chambers (9.5 cm diameter x 122-
cm length or 15 x 15 x 40 cm long) which were enclosed by a tempera-
ture-controlled water bath. Sinking rates shown in Table 3.2 vary from
0.01 to 0.83 cm -s"1, equivalent to the settling rates of inorganic sedimen-
tary particles of <1 to 100|iim equivalent diameter in size. Swimming
speeds in Table 3.2 vary from 0.01 to 0.50cm s"1, and if an average speed
of 0.2cm-s"1 was maintained throughout Id, the larva could swim a
distance of 173 m. Larval ciliary swimming speeds are strongly influenced
by environmental variables such as salinity and temperature because the
efficiency of ciliary propulsion is influenced by seawater viscosity; thus,
Hidu and Haskins (1978) were able to show that oyster larval swimming
speed was inversely related to temperature (Fig. 3.4).
Finally, the dispersion of larvae is critically dependent on ambient
currents, which are often of a larger magnitude than their swimming or
sinking velocities. The scales of relevance, in both time and space, are
Dispersal and Settlement 79
very broad and may include the effects of molecular and turbulent diffu-
sion, tides, storm-mixing events and wind-driven currents, Langmuir
circulation, internal waves, mesoscale eddies, and large-scale general
circulation (Okubo 1994).
A number of questions might be asked about the dispersion of larvae.
For instance, how are they distributed downstream of a point or area
from which they are released, either instantaneously or as a steady
stream, from the bottom of a benthic boundary layer? There is advection
of larvae along paths defined by stream lines, but they will also be subject
to turbulent diffusion, which will move them from the mean streamlines
in any direction. This problem can be analyzed in terms of the advection-
diffusion equation, an application of which will be discussed in Chapter
7. Its solution will give the concentration of larvae at any point down-
stream. This equation can be made to include terms for larval mortality
rate and for deposition on the bottom, if these are known.
In many circumstances, larval dispersion is dominated by more com-
plex flows than in the benthic boundary layer. For example, it is well
known that shoreward advection by currents driven by wind or waves
may be an important transport mechanism (LeFevre and Bourget 1992).
Two-layer flows in salt wedge estuaries and partially mixed estuaries
(Dyer 1979) will allow either landward or seaward transport of eggs or
larvae, depending on their depth. Other important flows which may
influence larval transport are tidally forced internal waves and bores, and
coastal up-welling, as will be subsequently discussed in more detail.
Topographically determined flows, such as observed around reefs and
embayments (Fischer et al. 1979), are not easily analyzed by use of the
advection-diffusion equation. Some of these topics are considered by
Okubo (1994). The physical oceanography of larval dispersal has also
been considered in conference proceedings edited by Sammarco and
Heron (1994).
An interaction involving one of the physical oceanographic processes
discussed and the swimming behavioral responses by the larvae of some
suspension feeders is important in determining the outcome of the dis-
persal process. We have collected the most prominent hypotheses con-
cerning such matters in Table 3.3.
Estuarine retention
In estuaries freshwater from a river is mixed with seawater from the
ocean by the action of tides, wind effects at the surface, and river dis-
80 Benthic Suspension Feeders and Flow
author showed that larval swimming depth increased with age so that
competent cyprids were found at the greatest depths. Position of larvae
within the water column varied with the tidal stage and current velocity.
Bousfield (1955) calculated that approximately 18 d later, -10% of those
initially spawned remained as competent larvae within the estuary, and
these were sufficient to maintain adult populations. By contrast, if the
larvae of Balanus improvisus maintained a uniform distribution in the
water column, his calculations showed that, as a result of seaward
advection, insufficient numbers would remain to recruit the historic
numbers of adult barnacles within this estuary.
In the study by Mann (1988) of the James River estuary, Chesapeake
Bay, in the United States, oyster larvae were reported to migrate through
a salinity discontinuity front present in the estuary and were then trans-
ported downwards, where they became entrained in the more saline
bottom water which moved landwards. Laboratory behavioral experi-
ments showed that intermediate stage oyster larvae could actively swim
through a salinity gradient (Mann 1988) involving a salinity difference of
3%o.
As pointed out by Stancyk and Feller (1986), some authors using
similar field observations consider that the observed larval distributions
support the null hypothesis of number 1, Table 3.3. In all cases, though,
the physical oceanographic conditions in the estuaries examined do not
conform to salt wedge or partially mixed conditions, rendering the estua-
rine retention hypothesis invalid in these conditions. Thus, DeWolf
Dispersal and Settlement 83
Shoreward advection
For open coastal environments such as wave pounded vertical rock faces,
the question as to how planktotrophic larvae manage to recruit in a
narrow zone on the shore has intrigued many investigators. Barnacle
larvae have been a favorite in such studies because they are
planktotrophic and feed only in the pre-competent stage. Balanus
crenatus, for example, may spend from 12 to 20 d in the seawater of the
Wadden Sea (DeWolf 1973) before settlement occurs. The question
addressed here is, How does the competent cyprid regain the intertidal
rock face required for the completion of the barnacle life cycle?
Working on the North Wales, United Kingdom, coast, Bennell (1981)
suggested that Semibalanus balanoides settled on scraped natural rock
surfaces which were cleaned daily when the wind direction was offshore
and the sea state calm. In contrast, Hawkins and Hartnoll (1982), using
similar field techniques but on the Isle of Man, United Kingdom, found
a positive relationship between S. balanoides settlement density and
timing with onshore winds (wind direction and velocity for the hour near
high water) during May 1979 (see number 2, Table 3.3). In a California
study by Shanks (1986), daily settlement, mainly of Chthalamus sp., was
not correlated with wind direction for a 4-mo period in 1983 or wind
direction at high water. Settlement rates of cyprids here were found not
to be significantly different on days with onshore versus those with
offshore winds.
In the study by Shanks (1986), three possible mechanisms were consid-
ered to explain how barnacle larvae reached suitable rock settlement
sites in the intertidal:
Cyprids swam ashore.
Cyprids were randomly dispersed and deposited on the shore.
Cyprids utilized onshore currents to transport them to shore.
The larval swimming capabilities were considered to be insufficient for
the purpose, since cyprids are commonly found in abundance to ~10km
away from the coast (Shanks 1986) and up to 100 km offshore on the
French Atlantic coast (LeFevre and Bourget 1991). The second of the
possibilities has been used as the null hypothesis for numbers 2-5 in
Table 3.3, although field evidence, e.g. the neustonic concentration of
barnacle cyprids of Verruca stroemia discovered by LeFevre and Bourget
(1991), suggests that, at least for some species, this is not true. The initial
hypothesis of shoreward transport of larvae by internal waves was devel-
oped for crab megalopae (Shanks 1983), and an integral part of the
Dispersal and Settlement 85
Rough Slick
^VVW ^^ ^ ^r- w ^ ^r- w v w w w v
/ t \
7
0.60 n
a:
UJ 0.20 -\
a:
0.00
Figure 3.6 Recruitment density as mean number of recruits per settling plate
(n = 18) against time for local barnacle settlement of B. glandula and
Chthalamalus sp. on the mid-California coast (Farrel et al. 1991). Wind speed is
shown (north > 0, south < 0). Bars at the top show when internal waves occurred.
Plant canopies
Physical oceanographic interactions with seagrass meadows are dis-
cussed in Chapter 8 in the section, Marine plant canopies: seagrass
meadows, and interactions with kelp forests in the section, Hydrodynam-
88 Benthic Suspension Feeders and Flow
ics. In general, they involve the development of a skimming flow over the
plant canopy, dependent on a sufficiently high flow speed and canopy
density, and within the canopy a reduction in speed, dependent on stem
density, which results in reduced shear stress and, therefore, causes
increased sedimentation. The original reason for studying seagrass
meadows was to find an explanation of the higher species diversity and
animal abundance commonly found there. Thus, passive larval entrain-
ment within the canopy could be an explanation of the increased abun-
dance, which is more fully documented later in this section, where other
hypotheses to explain it are also considered.
The hypothesis tested which relates to plant canopies (number 6,
Table 3.3) originated with Eckman (1983). In this study, a field experi-
ment was set up on the Washington coast of the United States. It in-
volved removing bulrush Scirpus americanus marsh plants in small plots
and then inserting simulated marsh plant stalks consisting of drinking
straws into the sediment. Recruitment of facultative deposit/suspension
feeders such as a spionid and sabellarid polychaete, but mainly of
meiofauna, was shown to have a significant positive correlation with the
numerical density of straws, in agreement with a hydrodynamic mecha-
nism of larval dispersal. Field observations on the North Carolina coast
in a seagrass meadow consisting of 80% Zostera marina and 20%
Halodule wrightii were conducted by Peterson (1986) to determine
whether larval supply rates explained the higher density and biomass of
hard clams within the meadow. The hard clam Mercenaria mercenaria
was sampled in 0.25-m2 cores with a sieving procedure that obtained only
clams of >5mm valve height. The size range of 5-25 mm sampled at
annual intervals was considered to be a good indication of larval recruit-
ment. The results, shown in Table 3.5, suggest that significantly more
larval (0-year) recruits do appear within the seagrass beds than on the
unvegetated sandflat. However, Peterson thought that passive hydrody-
namic trapping of larvae within the seagrass meadow was only one part
of the story since the older hard clams were present at a significantly
greater density within seagrass than on the unvegetated sandflat. The
numbers of older hard clams shown in Table 3.5 are higher than numbers
of 0-year recruits, in part because several years are accumulated in
calculating this mean. The larger differential within older hard clams
between seagrass meadow and unvegetated sandflat was thought by
Peterson (1986) to be due to differential survival within the two habitats,
with more efficient post-settlement predation likely to be the cause of the
Dispersal and Settlement 89
1980 1981
greater losses on the sandflat. That predation was more efficient on the
sandflat and inhibited by seagrass roots and rhizomes had been demon-
strated by Peterson (1982) and others.
In considering the second alternative hypothesis of number 6 in Table
3.3, Eckman (1987) mounted a field study in a Zostera marina meadow
on Long Island, United States, in which the recruitment of two
suspension-feeding bivalves - the bay scallop Argopecten irradians and
the common jingle Anomia simplex - were determined at four sites
within the local area. Both larval bivalves settle first on eelgrass blades
by means of byssus threads. By increasing the numbers of available
eelgrass blades with plastic mimics, recruitment as measured at approxi-
mately weekly intervals was dependent on the velocity of seawater
through the meadow. Thus, high density eelgrass meadows had fewer
recruits than those with fewer blades impeding the flow, as a result of
differences of volumetric flow which passed through the meadow.
In all the plant canopy studies so far reviewed, recruitment was meas-
ured and the likelihood of post-settlement mortality at each site not
independently assessed, and hence excluded as a possible confounding
factor in determining whether the initial settlement rate was indicated by
the recruitment rate. One such study in which settlement rates were
determined directly was reported by Wilson (1990). Larval settlement
rates were determined in settlement traps which were embedded in the
sediment and the larval catch counted daily. Thefieldwork,conducted in
90 Benthic Suspension Feeders and Flow
Post-larval dispersal
Post-larval stages of many invertebrates frequently reappear in the water
column after their settlement and metamorphosis. This was documented
in a review by Butman (1987), which lists 33 field observational studies
involving plankton tows, traps, or seawater pumping in which post-larvae
of bivalves, polychaetes, plus a few gastropods and meiofauna were
found in the samples.
A dispersal function for post-larval blue mussels Mytilus edulis in the
Wadden Sea was proposed by Maas Geesteranus (1942). The young
mussels were passively transported by currents until they detected
a suitable substratum for their growth. Byssal detachment and
reattachment of the young mussels could occur many times until a suit-
able substrate was found. Life cycle studies undertaken by Bayne
(1964a) of blue mussels in the Menai Straits, North Wales, referred to
earlier, suggest that the pediveliger larva first settles on afilamentousred
alga. After settlement and metamorphosis to the post-larval or planti-
grade stage, the young mussel may detach and be passively transported
by seawater currents. Bayne (1964a) referred to primary and secondary
settlement processes to describe this part of the blue mussel life history.
However, McGrath, King, and Gosling (1988), in studies on the rocky
shores of the west coast of Ireland, found that pediveligers of M. edulis
settled directly onto the adult mussel reef without first settling on a
filamentous red alga. Sigurdsson, Titman, and Davies (1976) presented
field and laboratory observations designed to elucidate the mechanisms
involved in post-larval dispersal in bivalves. They showed that slow
currents induced post-larval bivalves to produce a long byssus thread
which was functional in drifting dispersal. For example, juvenile tellinids
Abra alba of lmm valve height could be induced to produce a
mucopolysaccharide byssus thread which was 3 cm long and 2-4|iim in
diameter. The byssus thread increases the surface area of the juvenile
bivalve significantly. Because the settling velocity in creeping flow re-
gimes is proportional to surface area, we can expect the settling velocity
of the juvenile bivalve to be significantly less, thus enhancing dispersal.
The post-larval bivalve was able to detach from its byssus thread, upon
which it passively sank to the bottom. Sigurdsson et al. (1976) thought
that byssus drifting was of importance for many bivalve species of up to
2.5 mm in valve height, inclusive of protobranchs, e.g. Nucula tennis, and
many lamellibranchs, e.g. Mytilus edulis and Mya arenaria.
Most of the studies described are observational and designed to aid in
92 Benthic Suspension Feeders and Flow
Local hydrodynamics
The focus for the rest of this chapter shifts to a much smaller spatial scale
- <0.1km. Of most concern in benthic studies are flows near solid
boundaries, whether relatively flat bottoms or more irregular shapes
such as boulders and rocky surfaces. Except at very low velocities and at
very small distances from the surface, theseflowsare turbulent. Creeping
flows are seen very close to the solid boundaries, and, although these
regions are thin, they may be large by comparison to typical sizes of
larvae or eggs.
The hydrodynamically smooth benthic boundary layer is typically
found over soft sediments and other fine-grained bottom materials. This
flow is characterized by the following:
The viscous sublayer, which is a viscosity-dominated layer near the bottom
where the flow is mostly laminar
The logarithmic layer, which is turbulent and where the mean velocity varies
as the logarithm of the distance above the bottom
The outer layer, which is also turbulent but with turbulent energy decreasing
with height
The free stream flow, which drives the flow below it (Soulsby 1983)
94 Benthic Suspension Feeders and Flow
100:
0.01
0 2 4 6 8 10 12 14 16
HORIZONTAL FLOW SPEED, cm s"1
Figure 3.7 Turbulent velocity profiles at different shear stresses over a soft
sediment: a, smooth, /* = 0.60cm-s"1; b, rough, U* = 0.82cm-s"1; c, rough,
i7* = 0.98 cm -s"1. The curved region of line a represents the viscous sublayer. A
240-fxm pediveliger larva of the blue mussel is shown to scale (based on Butman
1986).
Larval settlement/recruitment
Before we begin an overview of the rapidly expanding body of work
which deals with the detailed hydrodynamics of settlement of competent
suspension-feeding larvae, we should recall the general ecological fea-
tures that are involved.
The life history of suspension feeders begins with spawning, and field
studies have suggested that the timing of this event is seasonally adjusted
for each species (Thorson 1950; Lacalli 1981; Hawkins and Hartnoll
1982). As an example, Lacalli (1981) showed that spawning periods in
Passamaquoddy Bay, within the Bay of Fundy, Canada, e.g. of polynoid
polychaetes (Fig. 3.8), are correlated with the timing of the spring diatom
bloom in Passamaquoddy Bay. The mechanism which allows this linkage
to be made was first suggested in the experimental work of Starr,
Himmelman, and Therriault (1990). In this work, two species were exam-
ined, one of which was a suspension feeder, Mytilus edulis. Adult blue
mussels were triggered to spawn by release of a chemically unidentified
ectocrine produced by phytoplankton cells found in the Gulf of
St. Lawrence, Canada. In the laboratory, phytoplankton such as
Skelatonema costatum, Phaeodactylum tricornutum, and Thallassosira
nordenskioldii produced sufficient ectocrine to induce mussel spawning.
96 Benthic Suspension Feeders and Flow
J FMAMJ J A S O N D
Polychaeta
Polydora quadrilobata
Scolelepis squamata
Harmothoe imbricata
Heteromastus filiformis
Capitella capitata
Nephtys incisa
Lepidonotus squamatus
Eteone longa
Prionospio steenstrupi
Nephtys caeca
Myriochele heeri
Pectinaria granulata
Phyllodoce mucosa
Phyllodoce maculata
Pholoe minuta
Nereis pelagica
Lumbrineris fragilis
Dorvillea caeca
Laonice cirrata
Other invertebrates
Boltenia ovifera
Barnacle nauplii
Strongylocentrotus droebachiensis
Cucumaria frondosa
Tonicella marmorea
Acmaea testudinalis
Asteroid larvae (Asterias sp.j
Tubellaria (Aiaurina sp.j
Echnarachinus parma
P i l i d i a (Cerebratulus lacteus)
Soft sediments
In a comprehensive review of invertebrate larval settlement in soft
sediments, Butman (1987) considered two major hypotheses - passive
deposition of competent larvae (number 1, Table 3.7) and active habitat
98 Benthic Suspension Feeders and Flow
with a microbial film but not processed by worms, or "tube" sand, pre-
pared from the broken off tubes that had been cemented by older
worms. Results showed that there was a unimodal response between
larval settlement rates determined in 3-h-long experiments and flume
bulk velocity flows. Peak settlement occurred at bulk velocities of 15-
25cm-s~1, and at these velocities, larvae tumbled end over end, some-
times briefly adhering to the floor by means of their tentacles. At
<10cm-s"1, larvae swam near the release point, eventually swimming to
the flume surface, where they were advected downstream in the bulk
layer, so that they did not come into contact with the settlement plate. At
>30cm-s~1, larval settlement was poor, either because the shear velocity
was too high and/or because the increased turbulence meant that fewer
larvae actually came into contact with the sand treatments. In field
conditions, Phragmatopoma is much more likely to experience oscilla-
tory wind-wave flows so it is not possible to place these results in a field
context.
Observations on the free-swimming larvae of the suspension-feeding
bivalve Ceras toderma edule in still water and a Vogel-LaBarbera flume,
3.5 m long x 0.5 wide x 0.4m high, were made by Jonsson, Andre, and
Lindegarth (1991). In still water, upward helical swimming of the compe-
tent pediveliger was observed and interpreted as geotaxis although
statocysts were not involved. In unidirectional flow flume observations,
four bulk velocities were offered: 2, 5,10, and IScm-s"1. At 5-10cm-s"1,
the larvae became confined to the viscous sublayer, where they drifted in
a streamwise direction at 0.45-1.60 mm-s"1, periodically touching the
flume floor. At 15 cm s"1, bedload transport of tumbling larvae occurred
with a much higher probability of larval resuspension. Jonsson et al.
(1991) suggested that boundary-shear forces created by t/>15cm-s" 1
prevented the larvae from swimming properly. An hypothesis involving
viscous drag torque on the larval body by benthic boundary layer forces,
which can help explain settlement by C. edule, is proposed as a specific
mechanism of viscous sublayer larval confinement by Jonsson et al.
(1991).
Spat of the Iceland scallop Chlamys islandica preferentially settles on
variousfilamentousred algae and on the perisarcs of dead hydroids in the
Gulf of St. Lawrence (Harvey, Bourget, and Miron 1993). Model studies
with three-dimensional branching structures made of plastic were de-
signed by Harvey et al. (1995) to determine the factors influencing larval
recruitment. The effect of branching patterns and branch diameter of the
red alga models on scallop larval settlement was determined in field
102 Benthic Suspension Feeders and Flow
treatment sites on a Puget Sound sandy beach. Each site had a core at the
centre, with either a bare sand control or a treatment site surrounded by
different densities of simulated marsh grass stalks. The latter were actu-
ally 0.6-cm-diameter plastic straws placed at three densities of 56-2500
straws m2. Coincident experiments in a recirculating flume where three
velocities, equivalent to U* = 0.52-1.67 cm-s"1, and similar straw densi-
ties to the field manipulations were set up. Enhanced recruitment in the
field was found for harpacticoids, amphipods, facultative deposit-sus-
pension feeding, tube-living polychaetes, and meiofauna. The results
closely follow those predicted from flume observations of boundary
shear stresses around rigid cylinder mimics and thus support Hx of hy-
pothesis 3 (Table 3.7). A similar type of field experiment with tube
mimics at densities less than that causing skimming flow showed that
enhancement of bacterial colonization occurred in the downstream
wakes (Eckman 1985). Bacterial colonization may be important for lar-
val recruitment since a bacterial film may stimulate larval settlement
(e.g., Crisp 1974).
Although the field experiments of Gallagher, Jumars, and Trueblood
(1983) were designed to test the well known conceptual model of
Connell and Slatyer (1977) of mechanisms of succession (inhibition,
tolerance, or facilitation), the results provide support for the passive
deposition hypothesis (number 3, Table 3.7). The experiments involved
placing cores containing clean sand into 10-cm2 patches to which live or
simulated tubes of the common tube builders in the Puget Sound, soft-
sediment intertidal area were added. Either live or simulated tubes en-
hanced the recruitment of other taxa to the de-faunated cores in com-
parison to that of controls without tubes, suggesting the physical nature
of the mechanism involved.
The hydrodynamic characteristics of the exhalant jet of a cockle from
Puget Sound, Clinocardium nutalli, were studied by Ertman and Jumars
(1988) in a 2.5-m-long-working-section recirculating flume. Exhalant
velocity in Clinocardium was estimated as 9-llcm-s 1 and patterns of
flow around it varied with ambient velocity in much the same way as
around the animal tube mimics studied by Eckman and No well (1984).
Polystyrene beads with the same settling velocity as that of narcotized
larvae were injected into the exhalant flow of an actively pumping cockle
and the sediment near it sampled with 0.6-cm-diameter plastic drinking
straws as corers. The free stream velocities tested ranged from 3.3 to
16.0cm-s"1, which was equivalent to [/* = 0.19-0.78cm-s"1. The results
were consistent with the passive deposition hypothesis (number 3, Table
106 Benthic Suspension Feeders and Flow
Hard substrates
Just as for soft sediments, for benthic ecologists interested in hard
substrates an important question is the prediction of the type of
macrofaunal assemblage which will develop on a particular de-faunated
area of rock or coral. Like those of larvae settling on soft sediments, the
ecological parameters of primary concern for hard substrate recruitment
are planktonic survivorship, larval settlement rate, and post-settlement
mortality. One way that larval settlement of suspension feeders of hard
substrates appears to differ from that of soft sediment is the absence of
post-settlement dispersal in the former. Hypotheses which have con-
cerned researchers in this field are similar to passive deposition (number
1, Table 3.8) and active habitat selection (numbers 2 and 3, Table 3.8)
of soft sediments. An additional hypothesis, which is more ecological
in outlook, has focused on larval supply and settlement rates, which
are proposed to control the final densities achieved (number 4, Table
3.8). This approach has been referred to as "supply side ecology"
(Roughgarden et al. 1987; Underwood and Fairweather 1989), although
many of the studies in support of it have not adequately separated larval
settlement rates from early post-settlement mortality rates.
Much of the field and laboratory experimental work on hard substrate
settlement has been done with the competent settling stage, or cyprid, of
cirripede barnacles. Crisp (1955) conducted settling experiments with
cyprids in 1.5-m-long, smooth-walled, glass tubes of from 2- to 10-mm
internal diameter. Water flow rates were varied by adjusting a constant-
head device and recording the volumetricflowrate. All of theflowswere
laminar. The velocity gradient at the wall of the tube can be related to
the average velocity and the tube diameter. The friction velocity, /*, is
related to the velocity gradient at the wall by
= E/*V1
dr
)o
Friction velocities have been calculated for the data of Crisp (1955); they
are shown in Table 3.9. The experiments involved introducing individual
Dispersal and Settlement 107
cyprids into the tube in a pipette and then observing their passage in a
1-m section of the tube. Results were expressed as a percentage of those
tested which made contact with the tube walls as a function of flow
velocity. Larval initial contacts were unimodally related to a range of
tube flow velocities tried, equivalent to velocity gradients up to 1000 s"1
for two species of barnacle, as shown in Table 3.9. Maximum attachment
was at velocity gradients up to, but less than, that in which the swimming
cyprid could maintain its position by swimming against the current. Peak
settling, b, occurred at -75 s"1 for Semibalanus balanoides and -40 s"1 for
Elminius modestus. In stage a response, the larvae did not swim and were
moved passively. Flows greater than 50 s"1 caused the cyprids to swim and
attach. Presumably in the c stage, increasing shear made it harder for the
cyprid to attach, until at >400cm"1 for S. balanoides and >700cm~1 for E.
108 Benthic Suspension Feeders and Flow
Faired Plate
Bluff Plate
Split Plate
c.
Figure 3.9 Streamlines over three types of settlement plate placed horizontal
with respect toflumeflow(Mullineaux and Butman 1991).
vertical advection
CO
< ."* "V
Q
first contact
settlement
erosion of the cyprids from the substrate because Eckman et al. (1990)
have shown that the critical drag force required to dislodge cyprids is not
reached at bulk flows of lOcm-s"1. The flume studies of Eckman et al.
(1990) showed that the force required to detach B. amphitrite cyprids was
positively correlated with the duration of attachment in periods of up to
3h.
Mullineaux and Garland (1993) reported on flume and field work with
Dispersal and Settlement 111
a
Refer to Table 3.8 for hypotheses tested.
Dispersal and Settlement 113
greater at the site occupied by the cyprid in comparison with vacant sites.
The results suggest that larvae can discriminate microheterogeneity
down to a length of 35|im, which is approximately the size of the
anntenular disc used by the cyprid both to explore and to attach to the
substrate. Again, it is not possible to decide from these field observations
whether the results are due to a passive or active process.
In common with earlier writers who used non-flowing larval bioassays
(e.g. Crisp and Meadows 1963), Rittschof, Sansford Branscomb, and
Costlow (1984) assumed that the chemical cues associated with settling
were adsorbed onto surfaces. The method they used to demonstrate
chemical attractiveness was similar to the glass tube bioassay for cyprid
settling in flow of Crisp (1955). B. amphitrite larvae were used in these
tests with a constant flow rate of 6.8ml-min"1 in a 3.25-mm-internal-
diameter tube, which gave a velocity gradient 0.5 mm from the wall of
39.4 s"1. The experiments carried out by Rittschof et al. (1984) showed
that rinsing the glass tube with 2.5 jug ml"1 of settlement factor, consisting
of ground-up, whole B. amphitrite and adjusted to 50|ig protein-ml"1,
significantly increased settlement in young cyprids compared to a non-
rinsed or control tube rinsed with bovine serum of the same protein
content. Besides attractant chemicals from conspecifics of poorly defined
identity, biofilms on surfaces have been identified as a settlement cue.
Thus, enhancement of settlement of B. amphitrite cyprids (Maki et al.
1990) or inhibition (Maki et al. 1988) by biofilms may occur.
Neal and Yule (1994) designed an experiment in which B. perforatus
and E. modestus cyprids, after initial contact on two contrasting biofilms
- high (83 s"1) and low (15 s"1) shear stress - were tested for the tenacity
of their attachment to each substrate. Thefilmswere developed on cover
slips influmeflowsover a 2-mo. period. The high shear stress biofilm was
thinner and denser than the other and resulted in increased cyprid tenac-
ity. The low shear stress substrate was relatively thicker and looser and
caused cyprids to attach with lower tenacity. In these experiments, it was
not clear how the benthic boundary layer flow would influence the re-
sults, although both cyprid species tested can obviously choose between
the substrates offered.
In combined flume and field experiments in the St. Lawrence estuary,
Miron et al. (1996) were able to examine both passive and active controls
of initial settlement of barnacle cyprids. To distinguish between hydro-
dynamic and chemical settlement cues, responses by Balanus crenatus
cyprids to either live or ceramic model adult barnacles were compared. It
was found that spat were absent on control model barnacle panels but
Dispersal and Settlement 115
Summary
From the foregoing overview of larval research of benthic suspension
feeders, we conclude that larvae, rather than post-larvae, are the most
118 Benthic Suspension Feeders and Flow
References
Abelson, A., D. Weihs, and Y. Loya. 1994. Hydrodynamic impediments to
settlement of marine propagules, and adhesive filament solutions. Limnol.
Oceanogr. 39: 164-169.
Ackerman, J. D., B. Sim, S. J. Nichols, and R. Claud. 1994. A review of the
early life history of zebra mussels (Dreissena polymorpha): comparisons
with marine bivalves. Can. J. Zool. 72: 1169-1179.
Bachelet, G., C. A. Butman, C. M. Webb, V. R. Starczak, and P. V. R.
Snelgrove. 1992. Non-selective settlement of Mercenaria mercenaria (L.)
larvae in short-term, still-water laboratory experiments. J. Exp. Mar. Biol.
Ecol. 161: 241-280.
Baggerman, B. 1953. Spatfall and transport of Cardium edule L. Arch.
Neerland. Zool. 10: 315-342.
Bayne, B. L. 1964a. Primary and secondary settlement in Mytilus edulis L.
(Mollusca). J. Anim. Ecol. 33: 513-523.
1964b. The responses of the larvae of Mytilus edulis L. to light and gravity.
Oikos 15: 162-174.
1965. Growth and delay of metamorphosis of the larvae of Mytilus edulis
(L.). Ophelia 2: 1-47.
1976. The biology of mussel larvae, p. 81-120. In B. L. Bayne (ed.) Marine
mussels: their ecology and physiology. Cambridge University Press,
Cambridge.
Bennell, S. J. 1981. Some observations on the littoral barnacle populations of
North Wales. Mar. Environ. Res. 5: 227-240.
Bertness, M. D., S. D. Gaines, E. G. Stephens, and P. O. Yund. 1992.
Components of recruitment in populations of the acorn barnacle
Semibalanus balanoides (Linnaeus). J. Exp. Mar. Biol. Ecol. 156: 199-215.
Boicourt, W. C. 1982. Estuarine larval retention mechanisms on two scales, p.
445-457. In V. S. Kennedy (ed.) Estuarine comparisons. Academic Press,
New York.
Bousfield, E. L. 1955. Ecological control of the occurrence of barnacles in the
Miramichi estuary. Bull. Nat. Mus. Can. 137: 69.
Burke, R. D. 1986. Pheromones and the gregarious settlement of marine
invertebrate larvae. Bull. Mar. Sci. 39: 323-331.
Dispersal and Settlement 121
Heron (ed.) Coastal and estuarine studies. Vol. 45. The bio-physics of
marine larval dispersal. American Geophysical Union, Washington, D.C.
Olafsson, E. B., C. H. Petersen, and W. G. Ambrose. 1994. Does recruitment
limitation structure populations and communities of macro-invertebrates
in marine soft sediments: the relative significance of pre- and post-
settlement processes. Oceanogr. Mar. Biol. Annu. Rev. 32: 65-109.
Pawlik, J. R. 1986. Chemical induction of larval settlement and metamorphosis
in the reef building tube worm Phragmatopoma californica (Polychaeta:
Sabellaridae). Mar. Biol. 91: 59-68.
1992. Chemical ecology of the settlement of benthic marine invertebrates.
Oceanogr. Mar. Biol. Annu. Rev. 30: 273-335.
Pawlik, J. R., and C. A. Butman. 1993. Settlement of a marine tube worm as a
function of current velocity: Interacting effects of hydrodynamics and
behaviour. Limnol. Oceanogr. 38: 1730-1740.
Pawlik, J. R., C. A. Butman, and V. R. Starczak. 1991. Hydrodynamic
facilitation of gregarious settlement of a reef-building tube worm. Science
251: 421-424.
Peterson, C. H. 1982. Clam predation by whelks {Busycon spp.): experimental
tests of the importance of prey sizes, prey density, and seagrass cover.
Mar. Biol. 66: 159-170.
1986. Enhancement of Mercenaria mercenaria densities in seagrass beds: is
pattern fixed during settlement season or altered by subsequent
differential survival? Limnol. Oceanogr. 31: 200-205.
Pineda, J. 1991. Predictable upwelling and the shoreward transport of
planktonic larvae by internal tidal bores. Science 253: 548-551.
Pratt, D. M. 1953. Abundance and growth of Venus mercenaria and
Callocardia morrhiana in relation to the character of bottom sediments. J.
Mar. Res. 12: 60-74.
Pyefinch, K. A. 1948. Methods of identification of the larvae of Balanus
balanoides (L.), B. crenatus Brug. and Verruca stoemia O. F. Muller. J.
Mar. Biol. Assoc. U.K. 27: 451-463.
Ricciardi, A., R. Serrouya, and F. G. Whoriskey. 1995. Aerial exposure
tolerance of zebra and quagga mussels (Bilvalvia: Dreissenidae):
implications for overland dispersal. Can. J. Fish. Aquat. Sci. 52: 470-
477.
Rittschof, D., E. Sansford Branscomb, and J. D. Costlow. 1984. Settlement and
behavior in relation to flow and surface in larval barnacles, Balanus
amphitrite Darwin. J. Exp. Mar. Biol. Ecol. 82: 131-146.
Roegner, C, C. Andre, M. Lindegarth, J. E. Eckman, and J. Grant. 1995.
Transport of recently settled soft-shell clams {Mya arenaria L.) in
laboratory flume flow. J. Exp. Mar. Biol. Ecol. 187: 13-26.
Roughgarden, J., S. D. Gaines, and S. W. Pacala. 1987. Supply side ecology:
the role of physical transport processes, p. 491-518. In J. H. R. Gee and
P. S. Giller (ed.) Organization of communities past and present. Blackwell
Scientific Press, Oxford.
Sammarco, P. W. 1991. Geographically specific recruitment and post-
settlement mortality as influences on coral communities: the cross
continental shelf transplant experiment. Limnol. Oceanogr. 36: 496-514.
Sammarco, P. W., and M. L. Heron (Editors). 1994. The bio-physics of marine
larval dispersal. American Geophysical Union, Washington, D.C.
Scheltema, R. S. 1974. Biological interactions determining larval settlement of
marine invertebrates. Thai. Jugosl. 10: 263-296.
126 Benthic Suspension Feeders and Flow
In this chapter, our concern is primarily with the first filtration stage of
feeding as defined later. In the next chapter, we deal with the second
stage of suspension feeding, which involves the mechanisms of particle
capture on the appropriate surface. Because these two stages of feeding
are the only ones directly affected byflow,they are the only ones consid-
ered in physiological detail in this book.
Background
In suspension-feeding animals, feeding and growth involves a series of
coordinated steps. Each of these steps is dependent on the preceding one
(Table 4.1). We will refer to suspension feeding as inclusive of the first
four steps shown and reserve the term filtration for the first two only.
Filtration or clearance rates (see Chap. 2, the section, Initial feeding
responses) need not, therefore, equal feeding rates if particle sorting
leads to rejection as pseudofaeces. Growth encompasses the remaining
catabolic and anabolic steps. Growth rates need not be directly related to
the feeding rate because of differential digestion of sestonic particles.
For example, in the giant scallop, some microalgae pass through the gut
undigested (Shumway et al. 1985b). Residence times of food in the gut
may be increased or decreased (Bricelj, Bass, and Lopez 1984) and thus
change the contact time between ingested food, and digestive and ab-
sorptive surfaces. Thus, filtration rate is not necessarily equal to feeding
rate and feeding rate need not always be directly proportional to the
growth rate.
Two major categories of suspension feeding benthic animals can be
recognized by the extent to which the filtration process is dependent on
external flow (Vogel 1981; LaBarbera 1984):
128
Flow and Physiology of Filtration 129
Stage
Stage number Description
Feeding 1 - Transport of seawater past the filtration surface
2 - Capture of seston at the filtration surface
3 - Transport of particles to the mouth involving
sorting and rejection, e.g. as pseudofaeces in
bivalves
4 - Ingestion
Assimilation 5 - Transport through the gut
6 - Uptake across specific gut surfaces
Growth 7 - Anabolic metabolism
8 - Catabolic metabolism
Elimination 9 - Waste solids voided as true faeces
10 - Metabolic wastes voided as urine
PASSIVE
a ! b i c id
ACTIVE-INHALANT ACTIVE-EXHALANT
NORMAL TO FLOW DOWNSTREAM IN FLOW
U
Figure 4.1 Effects of velocity, U, on filtration rate, R: a, increasing velocity
enhances filtration; b, optimum filtration rates; c, velocity inhibition of filtration;
d, filtration suppressed by velocity.
sion feeders rapidly to increase both the time spent filtering and the
volumetric throughput for filtration to take advantage of temporally
discrete pulses of good quality sestonic food.
For active suspension feeders, the filtration response also depends
critically on whether the exhalant siphon flow follows or is opposed to
the principal direction of ambientflow.Where the exhalant flow opposes
the ambientflow,the adverse pressure field causes a reduction in filtered
flow, as extraciliary pump work cannot be done to overcome the applied
external flow force.
Cnidarians
The sea pen Ptilosarcus gurneyi is a common cnidarian of soft, sandy-
mud sediments in Puget Sound and the San Juan Archipelago of the U.S.
West Coast (Best 1988). Its body consists of a central supporting pedun-
Flow and Physiology of Filtration 133
cle inserted in the sediment, a joint, and a rachis. The sea pen is oriented
so that its flat surface faces the major prevailingflow.A parallel series of
semi-circular leaves carries the filtering elements. These filtering ele-
ments are polyps with radiating tentacles bearing pinnules which are
located on the downstream part of each leaf. The rachis is flexible and
bends downstream as ambient velocities increase.
Volume flow rates through P. gurneyi have been determined by esti-
mating the rate of dye movement to determine velocity and a measure of
the rachis area from the dye-stained part of the sea pen (Best 1988).
Volumetric flow rates were related to the square of the height since the
filter area increased linearly with size. The maximum volumetric flow
through the rachis occurred at an ambient velocity which was a function
of sea pen size: thus 6.5-8.5 cm -s"1 for small, 12-14cm-s"1 for medium,
and 14-18 cm-s 1 for large sea pens.
Velocity measurements between the leaves and near the polyps made
with a thermistor flow probe showed that near the polyps velocities were
much reduced at the highest free-stream flows tested (25cm-s1) as a
result of drag near the filtration surfaces.
Filtration efficiency (RE; see Chap. 2) was examined in flume experi-
ments by taking seawater samples upstream and downstream of the
sea pen (with a suction sampler so that sampling was possibly not
isokinetic). The sample was examined in a particle counter and the
efficiency of capture of each particle size determined. Smaller sea
pens were significantly less efficient than medium or large individuals.
Free stream velocities ranging from 1.5 to 6.0 cm s"1 significantly reduced
RE.
Filtration and seston uptake rates, determined from independent esti-
mates of volume flow rate and filtering efficiency as outlined, were equal
in P. gurneyi and unimodal functions of velocity (Fig. 4.2). For the sea
pen, increasing velocity in low flow conditions enhances the filtration
seston uptake rate because of the increase in volumetricflowthrough the
filter. The plateau reached around 7 cm s"1 (Fig. 4.2) occurs as ambient
velocity just begins to deform the sea pen. This process eventually causes
both volume flows and filtering efficiency to decline in the c stage be-
cause pinnules and tentacles are swept back as the flow increases and,
consequently, less filtering surface is presented to theflow.For small sea
pens, the c stage begins at 13-16 cm-s"1 and, in medium and larger ones,
at 18-21 cm s"1. Filtration/seston uptake rate inhibition in P. gurneyi can
be seen to begin in Fig. 4.2 but, because the highest velocities tested were
only 25cm-s~1, the full unimodal function responses were not obtained.
134 Benthic Suspension Feeders and Flow
/ ^ ^ " ^ a Medium
105
Li x
C
< i
i
i
LU i Small
IO4: i
O
DC
<
CL-
10
0 2 4 6 8 1012 141618 2022
VELOCITY, cm -S" 1
Figure 4.2 Seston uptake rate in Ptilosarcus gurneyi as a function of ambient
velocity: open circles, retention efficiency at 30%; closed circles, at 20%. The
unimodal response limits are suggested (modified from Best 1988).
Echinoderms
The stalkless crinoid Oligometra serripinnia from the Great Barrier Reef
was used in experimental flume tests by Leonard, Strickler, and Holland
(1988). This crinoid lives at 10-15 cm above the sediment-water interface
136 Benthic Suspension Feeders and Flow
80 -
A
12 16 20 24 28 36 40
U, cm-s"1
B
8 12 16 20 24 28 32 36 40
(J, cm -s"1
Figure 4.3 Normalized feeding effectiveness with respect to ambient velocity:
, Acanthogorgia vegae; , Melithaea ochracea; and A, Subergorgia suberosa; A,
polyp feeding rates; B, feeding rate per unit surface area (Dai and Lin 1993).
Flow and Physiology of Filtration 137
50 - 50- B
LU 40 - 40-
O ai b i ,c
f
r/v
30 - 30
RCEI
20 - a |
; i
20
b i c
LU
Q_
10 - 10"
^ ^
2 4 6 8 10 12 14 2 4 6 8 10 12 14
50 - 50
aj b |
i
LU40 40-
| 30- 30-
O 20 20-
E
"10
2 4 6 8 10 12 14
VELOCITY C c m - s " 1 )
10-
(/. \
2 4 6 8 10 12 14
1
VELOCITY Ccm -s" 3
Bivalves
Although many early bivalve biologists (e.g. Kerswill 1949) recognized
the importance of flow in bivalve production, the precise mechanisms by
which this was achieved were not addressed until relatively recently.
Kirby-Smith (1972) was the first to study the effect of flow on bay
scallops (Argopecten irradians) in a growth tube apparatus (Chap. 2).
Flow was measured as the bulk velocity flow rate. Kirby-Smith (1972)
was also the first to demonstrate that higher velocities could inhibit bay
scallop growth. Wildish et al. (1987) have reinterpreted his growth re-
sults as a unimodal function of velocity, in general agreement with later
studies made in a flume with this bay scallop (Cahalan, Siddall, and
Luckenbach 1989; Eckman, Peterson, and Cahalan 1989). Growth re-
sults with the giant scallop {Placopecten magellanicus) in flume experi-
ments also show that the growth of this species has a unimodal function
response to velocity. Flume studies (Wildish and Saulnier 1993) where a
wide range of velocities were tested confirm that filtration by the giant
140 Benthic Suspension Feeders and Flow
0.2-
R
0.1-
Figure 4.5 The seston uptake rate, R, as arbitrary units, is a function of velocity,
U, in cm-s 4 , and seston concentration, C, as cell number-ml"1 of Chroomonas
salinus (Wildish et al. 1992).
valves would suggest that the pressure defect would be relatively small in
the vicinity of the exhalant, which will lie in the turbulent wake of the
scallop. The bulk stream velocity required to give a stagnation pressure
of 1.0 mm of water is 14cm-s"1. Because the flow through the trophic
fluid transport system is laminar, the volumetric flow is linearly related to
the external pressure difference (see J0rgensen et al. 1986a). If the
scallop's ciliary pump can generate a pressure difference of 1.0 mm of
water, the volumetric flow through the animal can be doubled by an
external pressure difference of the same magnitude. How the ciliary
pumping system of other bivalve molluscs would respond to extreme
pressure differences caused by ambient flow is not clear. As has been
demonstrated by Wildish and Saulnier (1993), by direct video viewing of
the exhalant siphon, the live scallop reacts behaviorally by partially
closing the mantle and valves in order to resist the external pressure
142 Benthic Suspension Feeders and Flow
Seston Seston
concentration Velocity Flux rate uptake rate
cells -mr 1 cm s"1 cells cm2 s"1 jag Chi a-g wt-h"1
10 50 500 0
100 5 500 0.018
50 5,000 0
1,000 5 5,000 0.09
50 50,000 0
10,000 5 50,000 0.30
50 500,000 0.24
100,000 5 500,000 15.0
Flow
Seawater velocity
direction
Mantle. Gills and Mouth Palps
photoreception
mechanoreception
chemosensory
Sensors 1
tactile
pressure
Integrated control
r
Nervous System
Hormones
i
Mantle Gills Mouth Palps Diaestive Svstem Growth
Responses opening pump on seston selection feeding rate positive
closing pump off seston rejection satiation negative
gill bypass to tissue
seston selection to gonad
to shell
10
me ifc sbc
-Vmn **&-
Figure 4.7 Diagrammatic representation of the blue mussel pump with typical
static pressures (J0rgensen et al. 1986a) at points along the trophicfluidtransport
system: A, relative pressure heads; B, diagram of pump; hx and h2 hydrostatic
pressures at inlet and outlet; A/Z12, difference between hx and h2 or back pressure;
in, and ex, inlet and exit losses of pressure; io, inhalant opening; me, mantle
cavity; If, laterofrontal cirri; p, pump; lc, lateral cilia; ifc, interfilament canal; f,
filament; sbc, suprabranchial cavity; es, exhalant siphon; 1, interfilament canal
width; vx and v2, velocities at inhalant and exhalant.
Bryozoans
The phylum Bryozoa consists of colonial forms, usually attached to the
substratum as an encrusting, stolonate, or bushy growth. The individual
zooid of the colony is protected by a non-living chitinous skeleton, some-
times strengthened by calcareous secretions. The characteristic tentacle-
bearing lophophore is the structure concerned with suspension feeding.
Observation of feeding Bryozoa in a static experimental system, where
the seawater was slowly mixed by a peristaltic pump, was undertaken by
Best and Thorpe (1983) in Flustrellidra hispida. This species is an epi-
phyte of the brown macroalga Fucus serratus, which is found in the lower
intertidal zone of the Isle of Man. The experimental method involved
viewing the lophophore at lOOOx magnification with a stereomicroscope
to observe the rate of movement of the seston particles (= Tetraselmis
suecica of ~10-|Lim diameter at all densities up to 2xlO 5 cells-ml"1).
Because the particles were moving quickly across the field of view, it was
necessary to tune an oscilloscope trace to the same speed so that from
the distance travelled (=fieldof view diameter), a particle speed could be
calculated. Results showed that starved Flustrellidra could increase the
seston particle speed if increased concentrations were supplied by up to
130%. Best and Thorpe (1983) noted that speeds were variable within
the lophophore and further investigation revealed that particles located
at the centre of the lophophore had the greatest speed. Rates of particle
Flow and Physiology of Filtration 149
(4-6 cm s"1) and fast (10-12 cm s"1). The results show an inverse relation
of feeding with velocity, and both large and small colonies were unable
to accumulate beads at higher rates. At low and moderate velocities,
zooids from larger colonies were more effective in feeding than small
colonies, perhaps because the larger number of zooids could better di-
vert seawater to the colony.
Subsequently, Okamura (1987, 1990) tested the interactive effects of
particle size and velocity on the feeding success of Californian Bryozoa.
She was able to show in Bugula stolonifera that medium and large
particles (14.6 1.0 and 19.1 1.1 Jim) were ingested at rates greater
than small particles (9.60.5jim). Flow influenced the particle size in-
gested by B. stolenifera so that this bryozoan fed preferentially on larger
particles at the highest flow tested (10-12cm-s~1). Particle size alone
affected the feeding success of Bugula neritana; the medium and larger
particles were preferred. The mechanism of selective feeding may in-
volve size-dependent particle behavior, active selection, or rejection in-
volving alternative feeding mechanisms such as tentacle flicking.
buccal siphon
A B atrium
atrial siphon
pharynx
stigmata
mucous strand
mucous rope
endostyle
mouth ;.;'./.;..;.;:..:;.
parapodia of
'^segments 14-16
twelfth segment. The ciliated dorsal groove and cuplike organ of the
thirteenth segment gather the mucus and roll it into a food ball. When
the ball is sufficiently large, the parapodia stops pumping and the food,
plus mucus, is transported anteriorly by cilia in the dorsal groove to the
mouth for ingestion.
The Chaetopterus variopedatus pump is a muscular one. Riisgard
(1989) has used his constant-level chamber to investigate these worms
from the Swedish west coast. The methods employed were similar to
those used with mussels (J0rgensen et al. 1986a) and ascidians (Riisgard
1988). Video recordings of Chaetopterus confirmed that the mechanism
involved a positive displacement pump and that parapod beating de-
creased with increasing back pressures. Imposed back pressures set up in
the constant-level chamber caused the worm to reverse itself in the tube.
The maximum pressure rise in the tube and the total head loss are
approximately twice those in bivalves and five times those in ascidians,
consistent with the filter area's being relatively smaller than in those
species. The pressure drop across the mucous net as estimated by the
Tamada-Fujikawa equation requires knowledge of mesh opening size
Flow and Physiology of Filtration 155
Barnacles
It was determined by Crisp and Southward (1961) that changes in veloc-
ity can alter the type of feeding behavior of a number of adult cirripede
156 Benthic Suspension Feeders and Flow
PASSIVE
ACTIVE FEEDING
FEEDING
C J
C>3 cm-s" 1 )
-\ I 1
forward stroke recovery stroke switch
Figure 4.11 Active and passive feeding in Semibalanus balanoides (Trager et al.
1990).
v.
Figure 4.12 Diagrammatic view of a model sponge in a flow where a benthic
boundary layer is present (Vogel and Bretz 1972).
Sponges
Model studies were undertaken by Vogel and Bretz (1972) in sponges,
which have many small incurrent openings through the body wall. These
openings connect to the spongocoel and an excurrent opening or
osculum which opens dorsally in the animal so that it is high in the
benthic boundary layer (Fig. 4.12). Besides an active flagellar pumping,
passive flow in the sponge is induced by the Bernoulli effect resulting
from the velocity and consequent pressure differences at the incurrent
and excurrent openings caused by their different positions within the
benthic boundary layer. Experiments with live erect sponges
Halichondria bowerbanki obtained from the Massachussets coast, were
made at ambient flows from 0 to lOcm-s"1 (Vogel 1974). Flow measure-
ments at 4 mm into the osculum and 3 cm in front of the sponge were
Flow and Physiology of Filtration 159
Lampshells
LaBarbera (1977) proposed that active suspension feeders such as
lampshells or brachiopods may use ambient flows to "augment or even
160 Benthic Suspension Feeders and Flow
supplant their active pump," as has been suggested for sponges. Model
studies showed that induction of water movement through the dead
valves of Terebratulina unquicula, Terebratalia transversa, and Laqueus
californianus did passively occur at ambient flows of 2-4 cm s"1. It is also
known that brachiopods may actively pump at zero flows. However, as
pointed out by LaBarbera (1977), there is no evidence available to show
that the filtration rate increases as ambient velocities are increased.
The bilateral symmetry of brachiopods (Chap. 6) with two incurrent
openings 180 apart allows them to utilize oscillating flows if the ante-
rior-posterior axis of the valves is aligned parallel to the current direc-
tion. There is also evidence that some lampshells can behaviorally
reorient toflowdirection by means of pedicel musculature (see Chap. 6).
LaBarbera (1977) proposed that passive flow augmentation through the
brachiopod lophophore resulted in increased feeding rates and/or in
increased availability of dissolved oxygen and removal of wastes. Vogel
and Bretz (1972) and LaBarbera (1977) suggested that their ambient
velocity passive enhancement theory could apply to other species, e.g.
scallops. An independently derived theory, the inhalant/exhalant pres-
sure field theory, was proposed by Wildish et al. (1987) to explain
behavioral orientation to flow direction by scallops, which might also
apply to brachiopods. Certainly, in future studies, both theories should
be considered in designing experiments and testing specific hypotheses.
According to the inhalant/exhalant pressure field theory applied to
brachiopods, the lampshells would be considered as active suspension
feeders, and the low exhalant flows measured by LaBarbera (1981) in
still water indicate that the ciliary pump is much less effective than in
bivalves. Exhalant velocities of Terebratulina unquicula, Terebratalia
transversa, and Laqueus californianus ranged from 0.2 to 1.2 cm-s"1 and
flow within the trophic fluid transport system was laminar. Referring to
Fig. 6.5, the following orientation mechanisms are possible:
A. When the exhalant flow directly opposes the ambient flow direction and
when the ambient flow pressure exceeds the ciliary pump pressure either
extra work must be done by the pump to overcome it or the valves partially
close (as in scallops) to reduce the differential.
B. With the exhalant downstream with respect to ambient flow direction and as
the flow builds up so that the external pressure tends to force flow across the
ciliary pump at a rate faster than that maintained by the cilia alone, then
either the valves partially close to reduce it or filtration retention efficiency
declines.
C. In this orientation the valves present the maximum amount of drag opposing
the flow.
Flow and Physiology of Filtration 161
Deposit-suspension feeders
Some species of benthic infauna have the ability to switch between
deposit and suspension feeding. Ambient flow velocity is the environ-
mental cue thought to trigger the change of behavior. Thus at low veloc-
ities, these animals feed on particles present on the sedimentary surface,
while at higher velocities, they switch to suspension feeding.
Spionids
The tube-building spionid polychaetes are characteristically able to
switch from deposit to suspension feeding. The switch seems to encom-
pass a range of velocities from <2 to 5 cm s"1 when a worm population is
sampled (Taghon, Nowell, and Jumars 1980; Taghon and Greene 1992).
For Pseudopolydora kempi japonica and Boccardia pugettensis, increas-
ing velocity causes an asymptotic rise in the number of worms which are
suspension feeding. Thus 50% of worms are suspension feeding at -1.7
for the former species and 4.5 cm-s"1 for the latter species (measured
0.3 cm above the bottom). Partly because a wide enough range of veloc-
ities was not used in the flume experiments of Taghon and Greene
(1992), their results do not provide sufficient detail to determine whether
a unimodal feeding response to velocity exists.
Miller, Bock, and Turner (1992) have extended observations to Spio
setosa in a flume oscillating flow. The sinusoidal flows produced in the
working section were characterized by the maximum free-stream veloc-
ity reached in each excursion. The percentage of worms suspension
feeding (that is palp coiling) in still water is zero, rising to -90% at
6.5 cm-s"1 and thereafter dropping off (maximum velocity tested of
-lOcm-s"1). The feeding response shown by populations of S. setosa
appears to be unimodal with respect to maximum oscillation velocity, but
the absence of higher velocity tests prevents drawing conclusions. In
another experiment, the percentage of Spio showing palp coiling was
observed at five maximum free stream velocities (6.8-30cm -s"1). In this
162 Benthic Suspension Feeders and Flow
Bivalves
Some species of the bivalve molluscan family Tellinacea, although
mainly sediment interface deposit feeders, may, under some environ-
mental conditions, subsurface deposit feed or suspension feed. Levinton
(1991) suggests the following feeding classification for tellinaceans:
Swash rider suspension feeders, intertidal species such as Donax which mi-
grate on sandy beaches to maintain moist and relatively unexposed sediments
Sandy bottom suspension feeders, intertidal and nearshore subtidal species
with relatively large ctenidia specialized for suspension feeding
Sandy-mud bottom switching feeders, intertidal and nearshore subtidal
sandy-mud species such as Macoma and Scrobicularia, some species of which
can switch between deposit and suspension feeding.
Of the latter group, Hughes (1969) studied the suspension feeding of
Scrobicularia plana by using the direct method of measuring pumping
rate in apparatus designed by Drinnan (1964). Filtration rates were
estimated independently by indirect measurement of seston depletion in
100-ml dishes containing one Scrobicularia. In these studies, no attempt
was made to investigate the effect of flow and the environmental factors
causing deposit to change to suspension feeding. In the study by
Levinton (1991), the feeding behavior of three Pacific species of Macoma
from the San Juan Archipelago - M. nasuta, M. secta, and M. unquinata
- was studied with respect to water flow and sediment transport. These
studies were done in an annular flume and showed that deposit feeding
was inhibited by increasing velocities (only three free-stream velocities,
-10, -25, and 33-36cm-s"1 were tested). This involved a decrease in
deposit feeding radius from the siphon hole because the inhalant was
withdrawn as velocity was increased. Levinton (1991) observed a loss of
control of the inhalant siphon (quivering/flopping over) at a free-stream
velocity between 10.5 and 12.0 cm -s"1, suggesting that drag on the siphon
is sufficient to overcome its normal functioning.
Flow and Physiology of Filtration 163
Summary
The following criteria can be used for assessing the completeness, or lack
thereof, of our knowledge of the physiological mechanisms of filtration
in benthic suspension feeders:
That velocity has been accurately measured near the appropriate filter or
inhalant opening of the trophicfluidtransport system over a sufficiently wide
range of ambient velocities and directions
That the volumetricflowwhich enters thefilteror inhalant is known
That the retention efficiency of thefilteris known.
Most suspension feeders have not been studied over a sufficiently wide
range of velocities to determine effects on filtration, feeding, or growth
and, consequently, conclusions about the general applicability of the
unimodal function model may be premature. Some hypotheses concern-
ing filtration of benthic macrofauna which require further experimental
testing are summarized in Table 4.5.
Of the few species that have been studied adequately with respect to
velocity, inclusive of passive (the sea pen Ptilosarcus querneyi,
gorgonians of the genus Pseudopterogorgia, the crinoids Oligometra
serripinnia and Antedon mediterrana) and active suspension feeders
(scallops, Argopecten irradians, Placopecten magellanicus, plus a few
species of Bryozoa), all appear to fit the unimodal model. The cause-
effect mechanisms at each response stage of the model are poorly under-
stood and require further experimental investigations. These studies
need access to controlled flows in flumes (see Chap. 2) with well-defined
boundary layer flows which simulate field flows as closely as possible. In
general, in stage a transport limitation of filtration can be expected, in
stage b this effect is absent and ambient velocity has no influence on
filtration, whereas in stage cfiltrationinhibition related to velocity can be
expected. In active suspension feeders, the filtration responses may also
be interactively linked to seston concentration and quality, making them
complex and often difficult to interpret.
As far as we are aware, there have been no studies which investigate
the effect of a wide range of flume-controlled ambient velocities on
filtration or feeding of sieving suspension feeders (Table 4.5, number 1),
and a unimodal response hypothesis probably applies to most groups
considered. The exact nature of the compound passive-active suspension
feeding group is still in doubt because, in order to establish firmly that
passive supply of seawater to the organism is of physiological conse-
quence, it must be demonstrated in live animals that enhanced filtration
associated with enhanced ambient velocity results in increased growth
164 Benthic Suspension Feeders and Flow
(Table 4.5, number 2). The critical experiments required to establish this
have not yet been carried out. The question of the energetic costs asso-
ciated with ciliary pumping in bivalves (number 3, Table 4.5) is periph-
eral to the arguments presented in this chapter, although we return to
this subject in relation to the evolution of the ciliary pump in Chapter 9
(the section, Ultimate questions).
Another interesting unanswered question concerns the ambient veloc-
ity passive enhancement and inhalant/exhalant pressure field theories
and their possible applicability to both brachiopods and scallops. As
proposed, the hypothesis (Table 4.5, number 4) does not appear to be
incisive in its ability to provide support for one or the other of the two
Flow and Physiology of Filtration 165
References
Bayne, B. L., and R. C. Newell. 1983. Physiological energetics of marine
molluscs, pp. 407-515. In K. M. Wilbur (ed.) The mollusca. Vol. 4.
Academic Press, New York.
Bayne, B. L., R. J. Thompson, and J. Widdows. 1976. Physiology I, pp. 121-
206. /n B. L Bayne (ed.) Marine mussels, their ecology and physiology.
Cambridge University Press, London.
Beninger, P. G. 1991. Structures and mechanisms of feeding in scallops:
paradigms and paradoxes, pp. 331-340. In S. E. Shumway and P. A.
Sandifer (ed.) World Aquaculture Society, Baton Rouge, La.
Best, B. A. 1988. Passive suspension feeding in a sea pen: effects of ambient
flow on volume flow rate andfilteringefficiency. Biol. Bull. 175: 332-342.
Best, M. A., and J. P. Thorpe. 1983. Effects of particle concentration on
clearance rate and feeding current velocity in the marine bryozoan,
Flustrellidra hispida. Mar. Biol. 77: 85-92.
1986. Effects of food particle concentration on feeding current velocity in six
species of marine Bryozoa. Mar. Biol. 93: 255-262.
Birbeck, T. H., J. G. McHenery, and A. S. Nottage. 1987. Inhibition of
filtration in bivalves by marine vibrios. Aquaculture 67: 247-248.
Birkbeck, T. H., and J. G. McHenery. 1982. Degradation of bacteria by
Mytilus edulis. Mar. Biol. 72: 7-15.
Bricelj, V. M., A. E. Bass, and G. R. Lopez. 1984. Absorption and gut passage
time of microalgae in a suspension feeder: an evaluation of the 51Cr:14C
twin tracer technique. Mar. Ecol. Prog. Ser. 17: 57-63.
Cahalan, J. A., S. E. Siddall, and M. W. Luckenbach. 1989. Effects of flow
velocity, food concentration and particle flux on growth rates of juvenile
bay scallops Argopecten irradians. J. Exp. Mar. Biol. Ecol. 129: 45-60.
Carlisle, D. B. 1979. Feeding mechanisms in tunicates. Sci. Ser. 103, Canada
Inland Waters Directorate.
Crisp, D. J., and A. J. Southward. 1961. Different types of cirral activity of
barnacles. Phil. Trans. R. Soc. (Ser. B) 243: 271-308.
Dai, C.-F., and M.-C. Lin. 1993. The effects of three gorgonians from southern
Taiwan. J. Exp. Mar. Biol. Ecol. 173: 57-69.
Drinnan, R. E. 1964. An apparatus for recording the water-pumping
behaviour of lamellebranchs. Neth. J. Sea Res. 2: 223-232.
Eckman, J. E., C. H. Peterson, and J. A. Cahalan. 1989. Effects of flow speed,
166 Benthic Suspension Feeders and Flow
1977. Current-induced flow through sponges in situ. Proc. Nat. Acad. Sci.,
U.S.A. 74: 2069-2071.
1978. Organisms that capture currents. Sci. Am. 239: 128-139.
1981. Life in moving fluids: the physical biology of flow. Willard Grant
Press, Boston.
Vogel, S., and W. L. Bretz. 1972. Interfacial organisms: passive ventilation in
the velocity gradients near surfaces. Science 175: 210-211.
Walne, P. R. 1972. The influence of current speed, body size and water
temperature on the filtration rate of five species of bivalves. J. Mar. Biol.
Assoc. U.K. 52: 345-374.
Ward, J. E., H. K. Cassel, and B. A. MacDonald. 1992. Chemoreception in the
sea scallop Placopecten magellanicus (Gmelin). I. Stimulatory effects of
phytoplankton metabolites on clearance and ingestion rates. J. Exp. Mar.
Biol. Ecol. 163: 235-250.
Ward, J. E., and N. M. Targett. 1989. Influence of marine microalgal
metabolites on the feeding behaviour of the blue mussel Mytilus edulis.
Mar. Biol. 101: 313-321.
Warner, G. F. 1977. On the shapes of passive suspension feeders, p. 567-576.
In B. F. Keegan, P. 6. Ceidigh, and P. J. S. Boaden (eds.) Biology of
benthic organisms. Pergamon Press, Oxford.
Warner, G. F., and J. D. Woodley. 1975. Suspension-feeding in the brittle star,
Ophiothrixfragilis. J. Mar. Biol. Assoc. U.K. 55: 199-210.
Wells, G. P., and R. P. Dales. 1951. Spontaneous activity patterns in animal
behaviour: the irrigation of the burrow in the polychaetes Chaetopterus
variopedatus Renier and Nereis diversicolor. J. Mar. Biol. Assoc. U.K. 29:
661-680.
Widdows, J., P. Fieth, and C. M. Worrall. 1979. Relationships between seston,
available food and feeding activity in the common mussel Mytilus edulis.
Mar. Biol. 50: 195-207.
Wildish, D. J., and D. D. Kristmanson. 1985. Control of suspension-feeding
bivalve production by current speed. Helgol. Wiss. Meeresunters. 39: 237-
243.
Wildish, D. J., D. D. Kristmanson, R. L. Hoar, A. M. DeCoste, S. D.
McCormick, and A. W. White. 1987. Giant scallop feeding and growth
responses to flow. J. Exp. Mar. Biol. Ecol. 113: 207-220.
Wildish, D. J., D. D. Kristmanson, and A. M. Saulnier. 1992. Interactive effect
of velocity and seston concentration on giant scallop feeding inhibition. J.
Exp. Mar. Biol. Ecol. 155: 161-168.
Wildish, D. J., and M. P. Miyares. 1990. Filtration rate of blue mussels as a
function of flow velocity: preliminary experiments. J. Exp. Mar. Biol.
Ecol. 142: 213-220.
Wildish, D. J., and A. M. Saulnier. 1992. The effect of velocity and flow
direction on the growth rate of juvenile and adult giant scallops. J. Exp.
Mar. Biol. Ecol. 155: 133-143.
1993. Hydrodynamic control of filtration in the giant scallop. J. Exp. Mar.
Biol. Ecol. 174: 65-82.
Winter, J. E. 1978. A review on the knowledge of suspension-feeding in
lamellibranchiate bivalves with special reference to artificial aquaculture
systems. Aquaculture 13: 1-33.
Young, C. M., and L. F. Braithwaite. 1980. Orientation and current-induced
flow in the stalked ascidian, Styela montereyensis. Biol. Bull. 159: 428-440.
5
Mechanisms of seston capture
Nature of seston
The type of particles which may be collected by suspension feeders are
highly variable and dependent on the local conditions. Non-viable parti-
cles of silt, clay, sand, and detritus may be processed for the attached
microbiota. Viable particles include bacteria, phytoplankton, inverte-
brate larvae, and eggs, many of which are weakly motile. Collectively,
this material is referred to as seston (Fig. 5.1); it ranges in size from less
than 1 \xm up to 1000 |iim or more.
In order to understand the mechanics of collection of sestonic particles
properly, their size, shape, density, and settling velocities should be
known. Shapes range from spherical to very complex. Inorganic particles
are of densities often well known from their mineral content, and settling
170
Mechanisms of Seston Capture 111
BIVALVES SESTON
1 m -T- 2 log cm
Inertial
forces
dom nate
ADULTS
1 cm 4 - 0
1 mm - 1 Macroplankton
JUVENILES
LARVAE
.EGGS. Microplankton
Viscous
forces -3 Nanoplankton_.
dominate
, Bacteria
1 pm - 4 Ultrananoplankton
---5
Virus
J6
that all particles which strike the collector are retained. In engineering
applications, an adhesion coefficient (zA) is defined to allow for any
resuspension of particles so that the overall collection efficiency is the
product of the adhesion coefficient and the engineering collection
efficiency.
The rate at which particles are collected is, therefore, given by
ECUAEA
Following Shimeta (1993) we will also use a clearance rate for a circular
cylinder defined in units of volume cleared per unit time per unit length
of cylinder.
The passive suspension feeder is often configured as a large number of
individual collecting elements in a supporting structure. A typical exam-
ple is the sea pen (see the section, Passive suspension feeders, Chap. 4).
Particle collection theory is mainly based on the analysis of capture by
isolated elements of simple geometry or by filter beds composed of layers
of cylinders or spheres. The preceding definitions were written with a
single solid obstacle such as a cylinder in mind, but with suitable reserva-
tions they can be applied to the whole structure, in which case the
collection efficiency represents the sum of the contributions of each of
the collecting elements in the entire organism. The analysis can also be
applied to a single collecting element. Each collecting element is charac-
terised by its own approach velocity and upstream concentration, and an
174 Benthic Suspension Feeders and Flow
Mechanics of capture
Suspension feeders must process large volumes of water to extract suffi-
cient seston as food. The food is in the form of particulate matter, usually
in low concentrations, often less than a few milligrams per liter (mg L"1),
and must be removed from the water being processed. The means by
which this is done by benthic suspension feeders are many and varied.
We have discussed the classification of the major methods in Chapter 4.
Classification may also be based on the mechanisms by which the parti-
cles are removed (LaBarbera 1984):
Biological, involving active response of the animal
Physical
Sieving
Of all the physical mechanisms, sieving is the most obvious and simple to
understand. Water containing seston is passed through a porous obstacle
and particles larger than the passages forfloware retained. Although this
is a simple concept, the reality is more complicated. Most suspension
feeders utilize seston of a wide range of sizes and while large particles
may be sieved, smaller particles will also be separated simultaneously at
lesser collection efficiencies by other mechanisms. There are many ex-
amples, however, of suspension feeders in which sieving is the dominant
mechanism. The use of mucous nets by ascidians and polychaetes has
been discussed in the previous chapter (see the section, Sieving suspen-
sion feeders). These nets are used within an animal, and the pressure
differences required to pump water through the mesh have been calcu-
lated and in some cases reconciled with the mesh sizes inferred from
particle retention efficiency measurements.
In some freshwater aquatic insect larvae, seston is collected by pre-
senting a filter net to a passing flow. Loudon (1990) and Loudon and
Alstad (1990) have discussed the mechanics of particle collection by
freshwater caddisfly larvae. The filter is external to the animal, and the
seston particles after capture are collected by the forelimbs for passing to
the mouth. These filtering nets are typically arrays of cylindrical obsta-
cles and separate particles both larger and smaller than the openings
between the elements of the filter. Silvester (1983) and Cheer and Koehl
Mechanisms of Seston Capture 111
Direct interception
Consider a particle of neutral buoyancy in a steady flow around a right
circular cylinder. This particle will move with the fluid if there are no
other forces acting on it. Its path will coincide with the fluid streamline it
was on upstream of the cylinder. It will collide with the cylinder if this
stream line comes within one particle radius of the surface of the cylin-
der. If the retention efficiency is unity, the particle will be captured. This
mode of capture is called direct interception.
Mechanisms of Seston Capture 179
FR = cUlcd(j-)=cUlc8 (5.3)
NR = ER (5.4)
180 Benthic Suspension Feeders and Flow
The distance, 8, will be dependent on the flow field around the cylinder.
If the spacing of the streamlines is assumed to be unchanged as the flow
passes around the cylinder, 8 can be replaced by the particle radius,
giving
This approach is used by Rubenstein and Koehl (1977) and Shimeta and
Jumars (1991) as a first approximation to the solution of the problem. In
creeping flow, however, the fluid velocity near the cylinder is less than
the upstream velocity. By the principle of conservation of mass, the
volumetric flow between any two streamlines must be constant every-
where in the flow field. It follows that the spacing between the stream
line to the stagnation point and the limiting stream line must increase in
the downstream direction since the average velocity in this stream tube
decreases.
A comment on the mechanics of steady flow around submerged cylin-
ders is needed at this point. The fluid flowing around the cylinder is
subject to forces which arise from the inertia of the fluid elements and
also as a result of the viscosity of the fluid, which resists any deforma-
tions. If the inertial forces are small with respect to the viscous forces, it
may be possible to ignore them and find a solution for the streamlines
dependent on the viscous properties of the fluid. The criterion for this
condition is that the "Reynolds number" of the flow is small. The
Reynolds number is defined as
E iA 'dA2
--- \t
and
FR = 2AFcUlcd2pd;x (5.7a)
where
AF = 0.5(2 -lnfRe,))"1 (5.7b)
Gravitational deposition
A particle which has a greater density than the surroundingfluidwill sink
at its terminal settling velocity, ws, and its trajectory will cross the fluid
streamlines, resulting in capture if the surface of the collector is reached.
The encounter rate for a horizontal cylinder is given (Shimeta and
Jumars 1991) by
FG = cwslc(dc+dp) (5.8)
Most particles of interest will settle at a Reynolds number considerably
smaller than 0.2, for which the settling velocity can be calculated from
Stokess law by
Mechanisms of Seston Capture 183
where
pp = density of the particle
p = density of the fluid
|i = absolute viscosity of the fluid
g = gravitational constant
given that the particle is spherical and of known density. The encounter
rate for particles small with respect to the obstacle is then approximately
proportional to the square of the particle diameter and is independent of
the ambient velocity. Equation (5.9) cannot be used if the density of the
particle is not precisely known and inaccuracies may arise if the particle
deviates from a spherical shape.
The engineering efficiency can be calculated as a fraction of the up-
stream flux of particles:
f
which can be considered as being equivalent to the efficiency for particles
small with respect to the obstacle, or as a measure of the drift of particles
across the fluid streamlines towards the upper surface of the obstacle.
Inertial impaction
When fluid containing particles approaches an obstacle, the streamlines
curve and particles of densities greater than the fluid will tend to move in
straight lines because of their inertia. These particles are more likely to
be captured. This mechanism of capture is called inertial impaction and
is shown schematically in Fig. 5.2B.
The importance of inertial impaction is determined by the stopping
distance of the particle in the fluid, which is defined as the distance the
particle, after release in the fluid at a certain velocity, will move before it
stops. The stopping distance of interest here is that for the approach
velocity of the fluid. It is given by
(5,2)
184 Benthic Suspension Feeders and Flow
(5.16)
3n\idp
where K is Boltzmann's constant, Tis temperature (Kelvin), and \i is the
absolute viscosity of the fluid.
One way of evaluating the relative importance of the diffusion mecha-
nism in particle collection is to determine the Peclet number
Pe = ^ (5.17)
^ , (5-19)
There is some uncertainty about the value of the coefficient (Pich 1966).
Shimeta (1993) gives the following expression for the encounter rate
for a cylinder in low Reynolds number flow:
This result predicts that the encounter rate due to Brownian diffusion
increases with the diffusion coefficient, as would be expected, but also
with the velocity, which is less obvious. The efficiency of collection is
inversely related to velocity, but since increased velocity delivers more
particles to the region of the cylinder the net effect is to increase the
encounter rate.
Electrostatic attraction
If there is an attractive electrostatic force between a particle and a
collecting surface, the particle will move across streamlines and is more
likely to be collected. This is an important mechanism in gas cleaning,
which is exploited in electrostatic precipitators and in fabric filtration.
Particles in seawater, however, have a "surface charge that is neutralized
locally by counterions that concentrate on the solution side to form an
electrical double layer" (Spielman 1977). The electrical fields extend to
only about 0.02 Jim from the surface, and it is reasonable to expect that in
Mechanisms of Seston Capture 187
F = 4cQqlc
E
(3^)
and the efficiency index by
4Qq
*B = ,- .^ (5-22)
These equations were derived for particles and collecting surfaces which
are smooth, in the sense of having irregularities much smaller than
0.02 |im, conditions which are not commonly encountered in benthic
suspension feeding. Although the equations may not be directly applica-
ble in many cases, the electrostatic mechanism may still be important in
the final stages of collection as the particle approaches the surface. How
this mechanism interacts with direct interception will be discussed in the
next section.
In freshwater applications, because of the much lower ionic strengths
involved, electrostatic attraction should not be dismissed without careful
consideration (Shimeta and Jumars, 1991). Gerritsen and Porter (1982)
showed that relative ingestion of three sizes of polystyrene spheres by
Daphnia magna could be modified if the surface charges on the particles
were neutralized or if a non-ionic surfactant were added to reduce
wettability.
Combined mechanisms
Each of the mechanisms of capture has been analyzed on the assumption
that the other mechanisms are of negligible importance. Particles which
are massive enough to separate by gravity, for example, will often also
have enough inertia to separate from curvilinear streamlines and so be
removed by inertial impaction simultaneously. In many, if not most,
188 Benthic Suspension Feeders and Flow
0.01 0.1 1 10
Dp, micrometres
Figure 5.3 Combined collection by direct interception and Brownian move-
ment. Clearance rates for a right circular cylinder in milliliters cleared per hour
per centimeter of length versus particle diameter. Velocity of approaching
stream: lower three curves, 0.01 cm-s"1; upper three curves, 1.0 cm-s"1, solid,
dotted, and dashed lines, cylinder diameters of 10, 50, and 100 (im; dp<0.1 dc\
Re c <1.0.
tive, equations (5.6) and (5.20) can be added to give the combined
encounter rate. Following the procedure of Shimeta (1993), this calcula-
tion has been done for cylinders of three diameters at velocities of 0.01
and 1.0cm-s"1 (Fig. 5.3). The Reynolds number of the flow around the
largest cylinder at the highest velocity is unity. The ordinate is the clear-
ance rate per unit of length of the cylinder, equivalent to the encounter
rate divided by concentration of particles and collector length. This
definition of clearance rate, based on a unit length of the collector, is
different from the clearance rate, R, defined in Chapter 2, which is based
on a single animal, or 5, the seston removed by a unit weight of a
suspension feeder. The diffusion coefficient has been calculated from
equation (5.15) at 20C. The curves have been terminated where the
particle diameter exceeds 10% of the collector diameter.
In the Brownian diffusion-dominated regime, the predicted clearance
190 Benthic Suspension Feeders and Flow
N
ad
where
Q = Hamaker constant
rc = radius of the collector
rp = radius of particle
The term AF describes the flow field around the collector, equation
(5.7b). The term Q is a constant of proportionality in Hamaker's equa-
tion for the force between a spherical particle and a much larger collec-
tor. It depends on the properties of the molecules of the two surfaces and
of the liquid medium (Spielman 1977). Monger and Landry (1991) have
used a value of 3.7 x 1CT14 erg for general lipid bilayer systems in
seawater. The adhesion number can be interpreted as the ratio of van der
Waals attraction to hydrodynamic forces.
Spielman and Fitzpatrick (1973) solved the equation for the limiting
particle trajectory around a cylinder in the presence of an attractive
force. By knowing this trajectory, the collection efficiency can be calcu-
Mechanisms of Seston Capture 191
O >-
LU o
ccz
Q LU
11
LU Q_
LLJLU
E LU
J. (5.23)
C/D
LU
10"
LU
o
0.01 cm S'1
5 1Q- E 5
o
10-
0.01 0.1 1 10
PARTICLE DIAMETER, micrometres
Figure 5.5 Clearance rates for right circular cylinders of diameter lOjLim (solid)
and 100|Lim (dotted) at velocities of 0.01 and 1.0 cm-s"1, attributable to direct
interception and electrostatic attraction. The dashed line is taken from Fig. 5.4
for direct interception, Brownian diffusion for U= lcm-s"1; dc = lOjim.
a
Symbols defined in text.
Mechanisms of Seston Capture 195
10
i A B
1 0n- -^3 O
10 1 -
3
A
10 - ' * ' I1
icr2-
CO
10"5-
104- 10 7 -
i 1 0 -6.
o iff9
I
o I
C/)
CO DC
E
o c 3 L 9...,
LU n
10-
D
LU irT3 -2
10 - *o ^^ 10 - r.-.-.-.-A-------"^7
[
'-''','
DC 10~5- 4 ,/. sA
LU
LUI 10" -
O 10~7- I I I o 10" 0- .1
6
I
1
I
10 50
I
<
cc VELOCITY, cm
LU 10 -i r 1 Clay
1 1 5 ( r c _ 5 ( )|im) 2 Silt
o 10 -
^^-^ L^-1 5(rc = 5 |im) 3 Sand
= 50 |im) 4( DMA
^ * *
could have been used, as has been shown by Shimeta (1993), who used
equation (5.7) rather than equation (5.5).
Other geometries
Encounter rate predictions for spherical collectors under low Reynolds
number conditions can be found in Shimeta and Jumars (1991). Cylin-
ders of non-circular cross sections are not amenable to simple analysis,
except in the case of gravitational collection, because of the lack of
hydrodynamic solutions for the flow around them, or because of their
complexity if they exist.
Assemblages of collectors
Many passive suspension feeders do not present single collecting ele-
ments to the flow but instead use arrays in a more or less parallel
configuration. For example, consider the pinnules on the tentacles of the
sea pen (Best 1988), the octocorals (Patterson 1991a), the comatulid
crinoids (Mayer 1979), and the stalkless crinoid Antedon mediterranea
(Leonard 1989). Active suspension feeders may also use similar arrays,
as is the case with bivalve molluscs. The first effect of introducing addi-
tional collecting elements is to modify the velocity field near the surface
of the collectors. The term AF, which was defined in equation (5.7a) for
the case of flow around an isolated circular cylinder, is modified for
arrays of cylinders (Spielman 1977). Streamlines will be compressed and
direct interception efficiency will be improved. The second effect is that
the bulkflowwill be influenced. In the case of passive collectors present-
ing an array of collecting elements to a stream, the flow between the
collecting elements will be reduced and more of the approaching fluid
will be diverted around the array if the elements become more closely
spaced.
Active suspension feeders will be required to supply pumping energy
to move fluid through the filtering apparatus at a rate proportional to the
superficial velocity and the area of collecting surface presented under
low Reynolds number conditions.
Approximate solutions to the problem of the filtering of aerosols by
beds of cylinders have been developed for low Reynolds number flows.
The approach has been to simplify the problem to one in which the
particles are assumed to be small with respect to the diameter of the
collecting cylinders so that the velocity distribution need only be known
close to the surface of the cylinder. The influence of the neighboring
cylinders is reflected in the value of the term AF introduced earlier for
isolated cylinderflowsin equation (5.7b). This term can be related to the
volume fraction of fibres in the filter bed. The equations developed for
Mechanisms of Seston Capture 199
capture by an isolated fibre can then be applied with the new value of AF
(Spielman 1977). The functional dependencies of encounter rates by the
aerosol mechanisms on parameters such as velocity, particle diameter,
and collector diameter are unchanged.
Table 5.3. Passive suspension feeders for which aerosol theory seston
encounter efficiency index calculations are available.
presented by Shimeta and Jumars (1991). They point out that LaBarbera
(1978) used an inappropriate way of calculating the efficiency of seston
capture by direct interception. Further analysis by LaBarbera (1984),
incorporating the hydrodynamic retardation model of Spielman (1977)
and allowance for surface forces in seston capture by the brittle star, was
also criticized by Shimeta and Jumars (1991) on several grounds. Among
those grounds was that the hydrodynamics of the flow near the collector
was not fully defined, that simple measures of mainstream flow are
insufficient to characterise the appropriate velocity to the collector, and
that streamlines around the collector depend on the Reynolds number,
which, in the case of the brittle star, was >1 and therefore strictly not
covered by aerosol theory.
Most of the seston encounter efficiency results obtained by authors
shown in Table 5.3 suggest that direct interception was the predominant
seston capture mechanism. Patterson (1984,1991a), for example, found
that the details of the capture process in Alcyonium siderium were more
complicated than might be visualised by a simple application of aerosol
mechanisms. Capture sites were not evenly distributed about the colony,
and the frequency and location of capture were influenced by the veloc-
ity and turbulence of the ambientflow.Patterson also observed that cysts
which were caught on the aboral side of the tentacle were usually lost
before transfer to the mouth, leading him to suggest that large particles
are more likely to be caught effectively by the tentacles on the down-
stream side of a polyp. The efficiency indices were calculated by using
the tentacle width as a characteristic dimension, typical velocities near
the tentacle, and a settling velocity of cysts of 0.08 cm s"1. The indices for
a velocity of 30cm-s~1 were as follows:
Gravitational deposition 0.003
Direct interception 0.67
Inertial impaction 4 x 10~5
Diffusive deposition 8 x 10"12
For a velocity of 3 cm- s"1, the direct interception index was unchanged,
the gravitational and diffusive deposition indices were increased by a
factor of 10, and the inertial impaction index was reduced by a factor of
10. The indices suggest that direct interception was the dominant mecha-
nism at velocities commonly encountered in nature. It should be empha-
sised, however, that some of the assumptions used in the development of
the indices are violated. The efficiency of collection of cysts was also
directly measured at flow velocities ranging from 2.7 to 19.8 cm-s"1 and
found to decrease with velocity, qualitatively consistent with a dominant
202 Benthic Suspension Feeders and Flow
CAPTURED CAPTURED
L = 331 201 L = 593 284
250 -
N = 452 N = 452
125
DC
LU
GO
AVAILABLE PLANKTON ^> AVAILABLE PLANKTON
400 - /_ = 713620 /_ = 2021 8 1 7 ,
A/ = 1017 N =1017
LU
DC
200 -
1500 3000
SIZE, Mm
Figure 5.7 Size distributions of particles captured by Alcyonium siderium and
of the available plankton: L, length 1 SD; plots on left, based on number of
items; plots on right, based on biomass (Sebens and Koehl 1984).
Mantle cavity
B
F LF L
Figure 5.8 A, schematic of water flow through Mytilus edulis; B, mean flow
streamlines through three gill filaments: F, frontal cilia (beating to the plane of
the diagram; LF, laterofrontal cirri (effective stroke to the left); L, lateral cilia
(effective stroke to the right) (Silvester and Sleigh 1984).
Bivalves
Although there is a long history of scientific study of the mechanism of
seston capture in suspension-feeding bivalve molluscs, there is still no
consensus on how it is achieved. The earliest view was that seston was
removed from inhalant water by passing it through sievelike structures
or gills. Mucus was thought to contribute to this mechanism by removing
the already captured seston and/or providing the actual sieving net;
Mechanisms of Seston Capture 205
hence, the mucociliary sieving theory. The ciliary pump in active suspen-
sion feeders consists of ciliary bands with seston collected by another set
of cilia beating in the direction of the mouth (J0rgensen et al. 1984). The
seston may be collected upstream of the ciliary pump, as in bivalves, or
downstream, as in sabellid polychaetes (J0rgensen et al. 1984). The
mechanisms by which sestonic particles leave the viscous stream through
pinnules and gill filaments to enter the surface flow to the mouth are still
a matter of contention. The trophic fluid transport system of the blue
mussel Mytilus edulis is shown schematically in Fig. 5.8A.
The sieving hypothesis generally considered the laterofrontal cirri the
location where sieving occurred (Fig. 5.8B). The Reynolds number of the
flow, based on the size of the laterofrontal cirri, is of the order of 10~4
(J0rgensen 1983). Each laterofrontal cirrus is in fact a compound struc-
ture consisting of 22-26 pairs of cilia arranged in two parallel but alter-
nating rows (Moore 1971; Owen 1974). The spacing of the cirri is about
2.0-3.5 |im, and the branched cilia are separated by about 0.6 |um.
Silvester and Sleigh (1984), Silvester (1988), Jones et al. (1992), and
J0rgensen et al. (1988) provide a discussion of the dimensions of the cirri
and the interfilamentary channel. The cirri are arranged in alternate rows
so that the branched cilia overlap slightly to form a barrier to the passage
of particles. The alternate rows have an effective stroke of 90 directed
towards the frontal cilia, one row being in position to screen water
coming through the ostium, while the other, a half cycle out of phase, is
in the recovery stroke (Owen 1974). The gills of Mytilus edulis have been
seen to remove particles larger than about 5|iim with close to 100%
efficiency. Particles of 1-2 jam are removed with about 50% efficiency
and particles smaller than l|um are removed with about 10% efficiency
(J0rgensen 1975).
J0rgensen (1975) was among the first to doubt the sieving hypothesis
in bivalve molluscs. He pointed out that observations on the dependence
of retention efficiency on particle size were not consistent with the mesh
openings provided by the laterofrontal cirri. In addition, the pressure
drop expected across the laterofrontal cirri in a position which blocked
the flow in the interfilamentary passage was too high to be supplied by
the ciliary pump (J0rgensen 1981a). Silvester and Sleigh (1984) chal-
lenged the pressure loss argument on two grounds: that the cilia formed
a more complicated organ than the series of parallel cylinders assumed
by J0rgensen (1975), leaving larger openings for the passage of inhalant
water, and that the velocities calculated by J0rgensen in the
interfilamentary canals were too high. Observed particle retentions were
206 Benthic Suspension Feeders and Flow
contention that most of the flow is not sieved by the cirri. When the
laterofrontal cirri are immobilized by serotonin, the retention efficiency
of algal particles smaller than 6|im is drastically decreased but that of 14-
lam particles is little changed. J0rgensen et al. (1988) suggest that the
laterofrontal cirri do not act as sieves of the through current, but rather
as modulators in the process of particle retention, improving its effi-
ciency for the smaller particles. Recent work by Nielsen et al. (1993) has
lent support to this view. Videographs of particle tracks in the vicinity of
an isolated gill filament showed a large fraction of sestonic particles
passing from the through to the frontal currents without any contact with
cilia. This diversion of particles was largely absent when the laterofrontal
cirri were immobilized. In bivalve taxa other than mussels, the
laterofrontal cilia may be absent or less well developed, and this is
considered by Riisgard (1988) and Nielson et al. (1993) to be an adapta-
tion concerned with improving the capture of smaller particles.
Summary
It is clear that it is usually possible to identify the dominant aerosol
mechanism(s) operating for a specified passive benthic suspension
feeder after a relatively cursory examination of the parameters of the
problem. The more ambitious aim of calculating the efficiency of collec-
tion or the encounter rate for a whole animal from observations of
animal morphology, fluid flow conditions, and type and quality of seston
collected has rarely been successfully completed in the same animal.
Although of practical interest, further work of the latter kind may not be
very helpful in advancing our understanding of the mechanisms of sus-
pension feeding, which are the focus of this chapter.
Some of the details of the mechanisms of suspension feeding have
been clarified by the use of the aerosol theory, since their introduction to
marine biology in 1977, but there have not been as many advances as
might have been reasonably expected. Some of the reasons for limited
success of the aerosol theory are described later. One limitation is that it
is only well developed for the case of capture by long circular cylinders
of smooth surfaces at low Rec. Although the low Rec condition is often
satisfied, the collecting elements are often not circular in cross section
and the surfaces may have roughnesses and protuberances. Nor are they
always oriented normal to theflowvector and they may beflexiblerather
than rigid. In addition, the theoretical efficiency predictions are devel-
oped on the assumption that the seston diameter is much smaller than
208 Benthic Suspension Feeders and Flow
the collector diameter. In most of the field and laboratory cases studied,
seston sizes were widely distributed and frequently overlapped the diam-
eter of the collector. This may not be a difficulty in identifying dominant
mechanisms, however, since it is inevitable that seston particles close in
size to that of the collector will be captured mainly by interception.
Real flow situations are seldom as simple as the idealized flows treated
by the theory. Nearby structures may alter the streamlines about the
collector, although this effect may be compensated for by the analyses
developed for fibre filter beds. Oscillating flows typical of wave environ-
ments or the stochastic flow caused by turbulence may alter the encoun-
ter rates and have not been fully studied. In passive suspension feeders,
the collecting elements may see only a portion of the flow which ap-
proaches the whole animal. Much of theflowmay be diverted around the
animal. Some will flow through but not encounter the filtering elements,
passing between the tentacles, for instance. This means that the flux of
seston based on the approaching flow overestimates what the encounter
rate, summed over all the collecting elements, actually is.
Passive suspension feeders have proved generally easier to study than
active suspension feeders. The following conclusions can be drawn:
Much of the work that has been reported on benthic suspension feeders up to
1995 had dealt with animals with primary collecting elements close to cylin-
drical in shape (though there are many exceptions) and with a flow around
them of Rec order of 1 or less. On the other hand, the preferred seston usually
is not significantly smaller (say by an order of magnitude) than the diameter
of the collectors so that this constraint on the theory is not satisfied.
Direct interception is an important mechanism for most non-motile organic
particles captured by passive suspension feeders, but if the modernized
theory of interception is applied, surface properties must be known. This
difficulty may be more real in theory than in practice in view of fact that
surface irregularities of typical collection elements are often larger than the
characteristic distance over which the surface forces are important (order of
0.02 urn). It is unfortunate that the theory for direct interception does not
reach an asymptotic state at low values of the adhesion number so there is no
limiting expression, as there is for the other extreme of very large intersurface
forces. The "classical" expression for collection efficiency, equation (5.7),
should give a reasonable estimate for many situations.
Gravitational separations may be important for certain particles which have
a significant sinking rate, but few reports are found in the literature.
Brownian diffusion is an important mechanism for collection of particles of
the size of the order of 1 jam, e.g. bacteria.
At the typical scales and Reynolds numbers encountered, inertial impaction
is probably not a significant mechanism for living organic seston. Inorganic
particles are often much more dense and, although not common in the litera-
ture reviewed, some inertial effects might be expected under favorable cir-
cumstances of particles with large stopping distances in rapid flows.
Mechanisms of Seston Capture 209
References
Best, B. A. 1988. Passive suspension feeding in a sea pen: effects of ambient
flow on volume flow rate and filtering efficiency. Biol. Bull. 175: 332-342.
Brun, R. J., W. Lewis, P. Perkins, and J. Serafini. 1955. Impingement of cloud
droplets on a cylinder and procedure for measuring liquid-water content
and droplet sizes in supercooled clouds by rotating multicylinder method.
Nat. Adv. Comm. Aero. Report 1215. NACA, Washington, D.C.
Byrne, M., and A. R. Fontaine. 1981. The feeding behaviour of Florometra
serratissima (Echinodermata: Crinoida). Can. J. Zool. 59: 11-18.
Cheer, A. Y. L., and M. A. R. Koehl. 1987. Paddles and rakes:fluidflow
through bristled appendages of small organisms. J. Theor. Biol. 129: 17-
39.
Crank, J. 1956. The mathematics of diffusion. Clarendon Press, Oxford.
Davies, C. N., and C. V. Peetz. 1955. Impingement of particles on a transverse
cylinder. Proc. R. Soc. A. 234: 269-295.
210 Benthic Suspension Feeders and Flow
species of northeast American bivalves. Mar. Ecol. Prog. Ser. 45: 217-
223.
Rubenstein, D. I., and M. A. R. Koehl. 1977. The mechanisms of filter feeding:
some theoretical considerations. Am. Nat. I l l : 981-994.
Schrijver, J. H. M., C. Vreeken, and J. A. Wesselingh. 1981. Deposition of
particles on a cylindrical collector. J. Coll. Interf. Sci. 81: 249-256.
Sebens, K. P., and M. A. R. Koehl. 1984. Predation on zooplankton by the
benthic anthozoans Alcyonium siderium (Alcyonacea) and Metridium
senile (Actiniaria) in the New England subtidal. Mar. Biol. 81: 255-271.
Shimeta, J. 1993. Diffusional encounter of submicrometer particles and small
cells by suspension feeders. Limnol. Oceanogr. 38: 456-465.
Shimeta, J., and P. A. Jumars. 1991. Physical mechanisms and rates of particle
capture by suspension feeders. Oceanogr. Mar. Biol. Ann. Rev. 29: 191
257.
Silvester, N. R. 1983. Some hydrodynamic aspects of filter feeding with
rectangular-mesh nets. J. Theor. Biol. 103: 265-286.
1988. Hydrodynamics of flow in Mytilus gills. J. Exp. Mar. Biol. Ecol. 120:
171-182.
Silvester, N. R., and M. A. Sleigh. 1984. Hydrodynamic aspects of particle
capture by Mytilus. J. Mar. Biol. Assoc. U.K. 65: 859-879.
Spielman, L. 1977. Particle capture from low-speed laminar flows. Annu. Rev.
Fluid Mech. 9: 297-319.
Spielman, L. A., and J. A. Fitzpatrick. 1973. Theory for particle collection
under London and gravity forces. J. Coll. Interf. Sci. 42: 607-623.
Trager, G. C, J.-S. Hwang, and J. R. Strickler. 1990. Barnacle suspension
feeding in variable flow. Mar. Biol. 105: 117-127.
Wildish, D. J., and D. D. Kristmanson. 1993. Hydrodynamic control of bivalve
filter feeders: a conceptual view, p. 299-324. In R. F. Dame (ed.) Bivalve
filter feeders in estuarine and coastal ecosystem processes. Springer-
Verlag, Berlin.
Yoshioka, N., and C. Kansoka. 1972. Collection efficiency of an aerosol by an
isolated cylinder: gravity and inertia predominant region. Chem. Eng.
(Japan) 36: 313-319.
6
Behavioral responses to flow
The life form of benthic suspension feeders includes both infauna and
epifauna, and as Table 6.1 suggests, the majority are epifauna. The
central problem for both sessile and tube-living epifauna is that they
must grow upwards into the benthic boundary layer to reach good qual-
ity, or higher quantities of, seston, both of which may be height or
velocity dependent (Muschenheim 1987a,b). Yet, these suspension feed-
ing epifauna may be restricted by the increasing velocity encountered
higher in the benthic boundary layer, because it increases drag and the
likelihood of dislodgement for them. By contrast, infauna are less sus-
ceptible to dislodgement, which does not occur unless a substantial part
of the sediment is swept away.
Behavioral responses to flow include rheotaxis, which is a directed
response to flow direction involving locomotion or muscular turning of
body parts, and rheokinesis, which is a non-directed response causing
random movement proportional to flow velocity. Rheotropisms are
debatably behavioral responses but fit the definition of behavior in
Carthy (1958); that is they involve flow-induced differential growth per-
ceived as twisting stresses by the axial support system, e.g. of gorgonian
corals which have semi-conductor properties that provide electrical
stimuli for skeletal secreting cells (Wainwright and Dillon 1969). The
result is greater skeletal growth on the stimulated side.
This chapter is organized around two general responses of suspension
feeders to hydrodynamics. The first involves only non-radially symmetri-
cal suspension feeders which respond behaviorally so as to position
themselves toflowin order to optimize filtration and, hence, growth. The
second is suspension feeding epifauna, which respond behaviorally to
flow in order to minimize lift and drag on their body parts. In addition to
this, we present a preliminary discussion of the phenomenon of aggrega-
213
214 Benthic Suspension Feeders and Flow
Flood
Figure 6.1 The clam My a arenaria: in life the valves are buried with only the tips
of the siphon reaching the sediment surface.
cal feeding structures, e.g. many sabellids, and so orientation with re-
spect to a particular current direction is not usually of concern. This is
not the case in bilaterally symmetrical infauna, e.g. bivalves, where indi-
vidual orientation to current direction may be of more importance.
It was noticed by Vincent, Desrosiers, and Gratton (1988) that the
deep-burrowing soft-shell clam Mya arenaria L. was oriented on tidal
mudflats in the Gulf of St. Lawrence such that the clam's sagittal plane
was perpendicular to the current direction (Fig. 6.1). Tidal currents over
the clam flats here were predominantly bidirectional, flood and ebb, and
it was hypothesized that this orientation was a means of optimizing
filtration. In any other orientation with respect to bidirectional currents,
some degree of exhalant mixing with the inhalant, and therefore seston
dilution, would occur, resulting in availability of a lower seston load for
filtration. No field evidence to support this view could be found by
Vincent et al. (1988) when individual orientation was related to growth
and reproductive measures in older clams, and it was suggested that this
216 Benthic Suspension Feeders and Flow
FLARE
HOOD
extensions (Thomas 1994). Flume studies showed that hoods and flares
resulted in a decrease of fluid exchange with seawater above the colony;
thus the extensions were not linked to possible food depletion above the
worms. In flume experiments hood orientation with respect to unidirec-
tionalflowsor complete removal of the hood showed that the hood acted
to reduce tentacle deflection at higher flows when present at its normal
field orientation. Since deflection of filtering elements by flow in other
suspension feeders (Chap. 4) results in a reduction in filtering efficiency,
it is reasonable to suppose that the extensions assist the worms to filter
feed at higher limiting velocities. Further experiments involving filtration
or growth are necessary to test this hypothesis experimentally.
Sessile epifauna
The coral sea fans (Gorgonacea) are frequently variable in growth form
with a particular morphology dependent on water movement. Theodor
(1963) reported water movement-related morphology in the Mediterra-
nean gorgonian Eunicella stricta. This coral has a concave form in strong
wave surges, a fan shape in moderate flows, and a whip form in lesser
flows. Leversee (1976) reported two growth forms of the sea whip
Leptogorgia virgulata in North Carolina waters: a bushy form where
directionally unpredictable, turbulent currents were present, and fan-
shaped colonies where predictable, bidirectional currents were found.
The Caribbean reef coral Agaricia agaricites, studied by Helmuth and
Sebens (1993), grew in three forms: bifacial, horizontal unifacial, and
vertical unifacial (Fig. 6.3). These growth forms were related to water
movement, which was inversely related to depth. Thus, the bifacial form
was found at the greatest depths, where the water movement was least,
and appears to be locally adapted by opposing rather than being parallel
to the flow, as in the unifacial forms. Other authors who have studied
corals report that the fan is generally opposed, or perpendicular to, the
major bidirectional flow occurring at a particular field location.
Leversee (1976) was the first to demonstrate that the orientation of
fan-shaped gorgonian corals with respect toflowdirection was important
in determining the rates of prey capture in this passive suspension feeder.
Experimental tests were made in a 60-L recirculating flume in which the
fan-shaped form of Leptogorgia virgulata was presented to the unidirec-
tional flow (at a velocity of 4.0 0.05 cm-s"1), either opposed or parallel
to it. Artemia nauplii at a concentration of 20 3 L"1 were used as food,
and the rate of depletion of the nauplii monitored over a 3-h period. Of
220 Benthic Suspension Feeders and Flow
FLOW
Figure 6.3 Colony form of Agaricia agaricities in the Caribbean (Helmuth and
Sebens 1993). Amorphous encrusting forms may also be present, but are not
shown in the diagram.
six colonies tested, four showed significantly greater feeding when op-
posed versus when parallel to the flow. Similar results were obtained by
Helmuth and Sebens (1993) with the scleractinian coral Agaricia
agaricites in unidirectional flume flows of 3 and 6cm-s"1 and food sup-
plied as Artemia cysts at 1000 cysts-L"1. Results suggest that at 3cm-s~1,
significantly more cysts were captured by the opposed coral versus that
with the fan parallel to the flow. Although more than twice the number
of Artemia cysts were captured by opposed compared to parallel corals at
a speed of 6cm-s~1, this difference was not statistically significant. A
review of the shapes produced by growth of some passive suspension
feeders in different locations is presented by Warner (1977).
Helmuth and Sebens (1993) compared the feeding rate of three
growth forms of Agaricia agaricites found in Caribbean waters (Fig. 6.3).
Their results reveal that for a given unit area of feeding surface of this
colonial coral, more polyps were present in unifacial (horizontal
12.0 3.3, vertical 13.7 3.2 calices cm"2) than in bifacial (8.1 2.0
calices cm"2) forms, and that this led to higher feeding rates when ex-
pressed either on a per polyp or unit area basis. We have reanalyzed their
results in terms of additional data (total number of cysts captured for the
whole colony) supplied by Helmuth (pers. commun. 1995). The addi-
tional results (Table 6.2) suggest that the bifacial upright colony feeds
best at lower ambient velocities with a nominal feeding peak -IScm-s"1,
compared to the vertical unifacial colonies, which feed best at higher
velocities and with a peak at >30cm*s~1. The latter results are consistent
with the known field distribution of the water movement-related
Behavioral Responses to Flow 221
A B C D
Figure 6.4 Orientation of a sea fan to ambient flow from left to right: A, fan
blade parallel to theflow;B, blade at a low angle; C, blade at a larger angle to the
flow; D, blade perpendicular to the flow; B and C, unequal twisting stresses,
indicated by filled arrows, are present (Wainwright and Dillon 1969).
Side Top
ess, both could traverse an arc as great as 120 to achieve their preferred
orientation. Although the mechanism was not studied by LaBarbera, he
suggested that the pedicle adjustor muscles were involved although the
sensory organs and nervous pathways were not identified. In L.
califbrnianus, movement of the pedicle was preceded by three to five
coughs (a rapid sharp closure followed by a full valve opening), whereas
the rheotactic behavior of T. unquicula did not include coughing but
rapid rotation through 60 to 90 with the valves closed or slightly
opened or with the valves open and a slower rotation, about 1 s per 30
through a final arc of 20 to 40. One other species, Hemithryris psittacea,
showed similar rheotactic behavior but could only adjust through 45,
Behavioral Responses to Flow 225
Semibalanus
balanoides =>
Elminius
modestus =>
Light
t
Balanus
improvisus =>
Balanus
amphitrite
Balanus
trigonus
=>
axis is aligned in the same direction as the flow. Thus, in B. trigonus the
cirral fan must swivel through only 90 (Fig. 6.6) to oppose flows from
either direction. For S. balanoides in oscillatingflows,the cirral fan must
be swivelled through 180 to oppose the flow when it approaches the
barnacle from the carina end. There appear to be no comparative growth
studies for B. trigonus at different orientations to prove the contention of
Ay ling (1976) that the preferred orientation toflowallows this barnacle
to minimize energetic costs involved in suspension feeding.
Otway and Underwood (1987), working with an Australian intertidal
barnacle, Tesseropora rosea, showed that its preferred orientation was
with the rostrum and cirral fan opposed to the main flow direction, as in
S. balanoides. Major flow in this habitat was the wave backwash. In
growth experiments on plates cut out of the solid rock, the plates were
rotated through 90 or 180 and compared to controls which were undis-
turbed or prepared exactly as the treatments. Results showed that signifi-
cant depressions of wet weight occurred in rotated plates in 15- to 19-mo-
long experiments, but there was no effect on the rate of shell growth,
mortality, or egg mass production.
Preliminary descriptions of the behavior of stalked barnacles Trilasmis
fissum hawaiense (family Poecilasmatidae), which are epizoic ecto-
commensals on the mouth parts of spiny lobsters from Hawaii, have been
given by Bowers (1968). The behavior includes movement of the
cirral net and other movements of the peduncle in response to water
movements.
Tube-living epifauna
Macrofauna which build tubes that project into the benthic boundary
layer are usually, although not exclusively, suspension feeders. On the
basis of their adult tube dimensions, H, height above the sediment-water
interface, and tube diameter, Z), opposing the majorflowvector, the ratio
H/D (using consistent units for H and D) can be used to characterize the
main types (Table 6.3) as follows:
In addition, some tubes have one, or both, open ends opposed to bidirec-
tional flow; the ratio H/D is not applicable to them. The spar buoy tube
is so named because it is articulated near its base and can passively move
Table 6.3. Tube height, H, cm, above the sediment-water interface and maximum diameter, Y>, cm.a
Tube dimensions
Species Higher taxa H D H/D Reference
Tube normal to flow
Lanice conchilega Polychaeta, Terebellidae 2.0 0.5 4 Carey (1983)
Pseudopolydora paucibranchiata Polychaeta, Spionidae 3.5 0.5 7 Levin (1982)
Streblospio benedicti Polychaeta, Spionidae 3.5 0.5 7 Levin (1982)
Eudistylia vancouveri Polychaeta, Sabellidae 12.5 1.7 7 Merz (1984)
Ampelisca abdita Amphipoda, Ampeliscidae 0.5-1.0 0.2-0.3 3-5 Mills (1967)
Ampelisca vadorum Amphipoda, Ampeliscidae 0.4-1.0 0.2-0.5 2-5 Mills (1967)
Haploops fundiensis Amphipoda, Ampeliscidae 1.4 0.5 3 Wildish and Lobsiger (1987)
Tube opposed to flow
Net-spinning caddis larvae Trichoptera, Hydropsychidae Edington (1968)
Ampithoe valida Amphipoda, Ampithoidea Borowosky (1983)
Tanais cavolinni Tanaidacea Borowosky (1983)
Truncated cone tube
Spio setosa Polychaeta, Spionidae 5.0 5.0 ~1 Muschenheim (1987a)
Fabricia limnicola Polychaeta, Sabellidae 0.6 0.4 ~1 Levin (1982)
Mesochaetopterus Sagittarius Polychaeta, Chaetopteridae 0.9 0.6 -1 Bailey-Brock (1979)
Phyllochaetopterus verrilli Polychaeta, Chaetopteridae 0.9 0.6 ~1 Bailey-Brock (1970)
Spar buoy tube
Potamilla neglecta Polychaeta, Sabellidae 1.5-2.5 0.08 19-31 Wildish and Lobsiger (1987)
Complex tube shape
Diopatra cuprea Polychaeta, Onuphidae 2-5 0.5-0.8 2-10 Myers (1972)
a
In some cases, tube dimensions were estimated from the original figures.
Behavioral Responses to Flow 229
Figure 6.7 Lanice conchilega tubes; flow patterns and sediment scouring near
the tube are based on time-lapse photography of egg white particles (Carey
1983); A, side elevation; B, from downstream; large arrow, flow direction.
substrate, the tube is eventually bent up at 90to the substrate so that the
finished tube is normal to ambientflows.Tube building occurs at the rate
of 3-4 mm-d"1 in laboratory conditions.
No laboratory flume studies have yet been made of ampeliscid
amphipods (Fig. 6.8), which have a tube with an elliptical cross section.
The study by Eckman and Nowell (1984) considered only rigid right
circular cylinders. Enequist (1949) and Mills (1967) have described
ampeliscid tube building and feeding, aided by observations in aquaria
Behavioral Responses to Flow 231
Antenna current
Opening Food
accumulates
Approximate_
Pleopod current
sediment level
Membranous
area
Figure 6.8 Ampelisca abdita amphipod: position adopted for feeding and com-
plete tube (Mills 1967).
bottom and is quite flexible. No flows are induced in the tube, and the
end inserted in the substrate is plugged with sediment.
During feeding, ampeliscid amphipods sit ventral side up at the top of
their tube (Fig. 6.8), with the body along the long axis of it, braced by the
posterior peraeopods and with the first two pairs of peraeopod dactyls
hooked over the tube lip. If the animal moves lower into the tube, the lips
close after it, thus preventing outside access. Three feeding methods are
possible (Enequist 1949; Mills 1967): passive filtration, holding long
setose first and second antennae into the flow; whirling, movements of
both second antennae beating in synchrony, which creates a flow across
the gnathopods where particles are filtered, with the feeding flow aug-
mented by pleopod beating, which also drives the flow from anterior to
posterior; and deposit feeding, which is possible when the long second
antennae are scraped across the sediment surface. The specificity of
these feeding methods is not clear, although Mills (1967) considers A
vadorum and A. abdita as mainly "whirling" feeders, whereas Haploops
fundiensis must be a passive filterer or whirler since the second antennae
are not long enough to reach the sediment for deposit feeding from its
position in its tube.
a
Velocities are measured at palp height.
Source: Data from Taghon et al. (1980).
opposed to the flow. The time spent suspension feeding was related to
ambient velocity so that it predominated at maximum flows >15cm-s"1.
Deposit feeding predominated at maximum flows <10cm*s~1, and a
change to suspension feeding appeared to be related to the inability to
keep the palp tip on the sediment surface at velocities >10cm-s~1. The
relevant hydrodynamic variable to characterize the transition could be
the [bed] sediment shear stress. In Barnes's (1964) study of the same
species, the worms also spent time building their tubes, which involved
withdrawing their palps and exposing the fourth parapodia for this pur-
pose or tube cutting as velocities increased by means of the fourth setae.
Tube building in Spiochaetopterus was maximal in motionless seawater,
and within 15 min of initiating oscillating flow, tube cutting also started
(Barnes 1964; Turner and Miller 1991b). Reducing the height of the tube
is presumably related to adjusting the worm's feeding ambit to one where
good quality seston is available.
(Myers 1972). Its tube extends 2 (1-5 range) cm above the sediment
surface, depending on local sediment erosion/deposition near the tube.
Diopatra can reduce the tube height by cutting if erosion is marked and
build the tube upwards if considerable deposition of sediment occurs.
Where currents are high, greater worm densities are present and the
worms can orient the tubes behaviorally to within a 90 arc, so that
the long axis of the J, seen from above, is at right angles to theflow.The
worms may also rebuild the tubes so that a better orientation to flow is
obtained (Myers 1972). Although it was suspected that the tube orienta-
tion was connected with feeding, no experimental results confirming this
were presented.
A flume study of Diopatra cuprea was undertaken by Luckenbach
(1986) to determine the erodibility of sediments affected by tubes in flow
or by associated deposit-feeding macrofauna. During these experiments,
it was not stated how the tubes were oriented with respect to the unidi-
rectional flume flow. Those sediments with high macrofaunal densities,
with or without Diopatra tubes, were more erodible than consolidated
sediments with fewer macrofauna, suggesting that the tubes were unim-
portant in sediment stabilization, at least at friction velocity values em-
ployed during these experiments (Luckenbach 1986).
Free-living epifauna
The scallops (Pectinidae) are mobile epifauna with some swimming ca-
pabilities, which are more fully discussed later in the section, Swimming.
Because their body plan is bilaterally symmetrical with inhalant and
exhalant openings at specific positions on the body (Fig. 6.9) - as in the
brachiopods observed earlier - it is plausible to propose that rheotaxis
can optimize their orientation with respect to ambient flows.
Observations by SCUBA of natural populations of Pecten maximus in
Strongford Loch, Northern Ireland (Hartnoll 1967), showed that the
majority of scallops opposed the predominantly unidirectional flow
(59% of individuals at 45 to 135; see Fig. 6.9C). Mathers (1976) also
observed P. maximus on the west coast of Ireland, at two locations where
the tidal flows reversed on flood and ebb tides. Here it was found that
approximately half of the scallops opposed the ebb and half opposed the
flood currents with the ventral margin. The digestive histology of flood
and ebb subpopulations showed that digestion was offset by 6h, suggest-
ing that each group fed on either the flood or ebb tide only, but not on
236 Benthic Suspension Feeders and Flow
resillum
ANTERIOR
270 c
Figure 6.9 Scallop valve orientation and morphology: A, left upper valve from
above; B, right, lower valve, inside view; small arrows, inhalant; large arrow,
exhalant; C, valve orientation with respect toflowdirection (large arrow).
Table 6.5. Giant scallop seston uptake rates, jigChla -gwet^h1, mean
of two replicate experiments at each treatment?
Velocity, cm-s 1
Scallop 10 18 31
orientation
(see Fig. 6.9C) ng-g'-h-1 % ng-g'-h-1 % Hg-g'-h-
1
%
Speed (cm s'1) Scallop orientation () Valve height (h) Wet weight (W)
a
Conditions: 28-d duration starting on November 5, 1986, mean
temperature = 7.6C with range of 6.1-9.5; mean Chla = 0.65 (range 0.48-
0.99) ug/L, 30 juvenile giant scallops per treatment.
b
100 (hx - hoyhoN, 100 (Wt - W0)/WyV, where N = time in days.
Source: Wildish and Saulnier (1992).
limited when the exhalant faces the flow (225), but this effect is absent
at low ambient velocities (3.6 cm-s"1) for the same orientation.
Behavioral cues which must be involved in scallop orientation
behavior are not well known, although rheotaxis may be involved. The
ability of scallops to spin around by directing a jet of water from only one
side of the animal (Caddy 1968) may help to explain how it is done. An
alternative hypothesis that scallop orientation toflowis a passive process
238 Benthic Suspension Feeders and Flow
Sessile epifauna
The greatest hydrodynamic forces exerted on macrobenthos occur on
intertidal rocky shores. The reef coral Acropora reticulata was shown by
Vosburgh (1977) to be limited to depths >7.7m in Eniwetak Atoll, where
it escaped the most violent breaking waves. Drag and mechanical tests
Behavioral Responses to Flow 239
Species Re cD cL CM
5
Semibalanus cariosus 10 0.52 N/A 1.31
Mytilus californianus - end-on 105 0.20 1.20
- broadside 105 0.80 2.00
Models
Sphere 105 0.47 0.4 1.67
Cylinder 105 0.73 N/A 2.00
a
CD = drag coefficient, CL = lift coefficient, CM = inertia coefficient. (See the
section, Free-living epifauna).
Source: Denny (1985).
oriented with respect to the flow. On the wave-swept rocky shore, flow
directions are unpredictable and therefore streamlining is probably not a
realistic option. As Denny (1985) points out, mussels may be present in
densely packed reefs where the forces are shared by all rather than by an
individual mussel. Some intertidal organisms, e.g. Semibalanus cariosus,
arefirmlycemented to the substrate by a special adhesive to resist the lift
and drag forces involved.
Adhesive forces involved in Semibalanus balanoides attachment have
been studied by Yule and Crisp (1983) and Yule and Walker (1987) by
using strain gauges. Eckman, Savidge, and Gross (1990) used an indirect
method of determining drag and detachment forces for Balanus
amphitrite from shear forces in a small flume. The latter study suggests
that the drag and detachment forces associated with Balanus are lower
than those determined directly with strain gauges in Semibalanus.
Whether this is a real difference consistent with the more exposed life-
style of Semibalanus or is due to differences in measurement technique
awaits further standardization of both methods.
In terms of the preceding introduction to forces commonly present on
wave-swept rocky shores, we must ask whether behavioral responses to
flow are involved in helping the organism to avoid flow dislodgement.
Adhesive attachment of benthic organisms on a wave-swept rocky shore
is a common solution to dislodgement, e.g. in barnacles (see earlier
discussion), or mussel byssus threads which are glued to the substrate
(Waite 1983). Skeletal stiffening by spicule formation, e.g. in sponges
and cnidarians (Koehl 1982), and strengthening by coral formation or
development of various forms of body elasticity which allow form
changes proportional to velocity, resulted in reduced drag (Koehl
1911 di). Most of these must be considered as structural adaptations to
resist the forces generated by flow.
Size may also represent a refuge from the worst effects of orbital
wind-wave action, e.g. for the sea anemone Metridium senile (Anthony
and Svane 1994). Large numbers of small sea anemones of pedal disc
diameter of 0.3 cm are found where wave exposure is strong versus
fewer, much larger, 2.5- to 3.5-cm-pedal-disc-diameter individuals in
subtidal lower wave-exposed habitats.
Behavioral responses may be involved (Koehl 1984) in withstanding
moving water; they include adjusting the orientation of the individual to
minimize lift and drag. One of the characteristics shown by the sea
anemone Anthopleura xanthogrammica which live in exposed intertidal
surge channels of rocky shores on the U.S. West Coast was found by
Behavioral Responses to Flow 241
Free-living epifauna
Arnold and Weihs (1978) developed a hydrodynamic model for plaice
rheotaxis which is applicable to most benthic macrofauna where
Re = 10,000 - 1,000,000 (where Re is defined by multiplying body length
times free-stream velocity divided by the kinematic viscosity of sea-
water) as a means of analyzing the effect of flow forces in dislodging
free-living epifauna. In the model, the fish is assumed to be dislodged by
the flow when the drag force on it is just balanced by the static friction
force between the fish and the bottom. The static friction force is propor-
tional to the difference between the fish's buoyant weight and the lift
force generated by the flow over the fish.
An experimental study of mussel dislodgement of both Mytilus
californianus and M. edulis was made by Witman and Suchanek (1984),
using a direct drag force measurement system. The force that was re-
quired to dislodge the mussel by breaking the bysuss threads was re-
corded. It was found that attachment strengths increased with shell area
242 Benthic Suspension Feeders and Flow
and varied with position in the mussel reef and with the degree of wave
exposure of the habitat. The measured attachment strength of the byssus
threads was found to be proportional to the degree of wave exposure of
the mussel's habitat. The Re under flume flow conditions and based on
the valve length of the mussels used by Witman and Suchanek (1984)
ranged from 9000 to 21,000, but with simulated epifloral attachments of
considerable length, this increased to 53,000-27,000,000. Thus, mussels
that had been overgrown by kelp experienced flow-induced drag forces
that were two to six times greater than on the mussels alone, and this
explains why, during storms, mussels could be dislodged. Laboratory
experiments by Young (1985) were designed to determine the effect of
various environmental variables on the rate of byssus formation by M.
edulis. The most important factor was found to be periodic agitation,
simulating wave surges, which caused 15.8 threads d"1 for each mussel to
be produced. This rate was twice that of any other factor tested such as
temperature, salinity, tidal submergence, or season, although all of those
affected byssus thread formation. In agreement with results of Witman
and Suchanek (1984), Young (1985) found that wave-exposed mussels
had more byssus threads (87 per mussel) than those from sheltered sites
(48 per mussel).
Preliminary studies using the Arnold and Weihs (1978) model with
giant scallops Placopecten magellanicus have been presented by Wildish
and Saulnier (1993). In these experiments, the scallops were placed on
the smooth bottom of a flume and the velocities at which they were first
dislodged observed.
In order to study the hydrodynamics of the scallop, it is necessary to
study their body structures. The geometry of giant scallop valves (Fig.
6.10) shows that the valve length is less than the valve height at <10cm,
but is greater at >10-cm valve height. In natural conditions, the giant
scallop opposes the flow with its ventral edge. The frontal area Af pre-
sented to this flow (Fig. 6.10) is irregularly ellipsoidal in cross section,
and the lower, right valve is flatter than the upper, left valve. Because of
this irregularity, Af was determined empirically by Wildish and Saulnier
(1993) from photographic cutouts calibrated to area (rather than by
calculation of n\w). Seen from the posterior edge (Fig. 6.10), the maxi-
mum valve width does not occur centrally but more dorsally as shown.
Because of the choice by the scallop to oppose the flow with the ventral
edge first, i.e. with the "trailing edge" (Fig. 6.10) rather than the "blunt"
end, it is not optimally streamlined for lift (like the aeroplane wing). The
potential hydrodynamic instability of this orientation has been discussed
Behavioral Responses to Flow 243
anterior
Figure 6.10 Geometry of the adult giant scallop valves showing the measuring
origins adopted: h, valve height; /, valve length; w, maximum valve width or
thickness; g, maximum valve gape.
viscous and form drag on the valves will exceed the restraining force
exerted on the left valve by the sediment, and the scallop will slip down-
stream. At this slip speed, Arnold and Weihs (1978) show
FD=K(W-FL) (6.1)
where FD is the drag force, K is the static friction coefficient, and W is the
submerged weight of the scallop.
The lift is defined as
FL = 0.5psAfCLU2 (6.2)
D = 0.5psAfCDU2 (6.3)
- ^ - T = CL+-^- (6.4)
?AfU] %
The term on the left of the equation where Us is the slip speed, cm-s"1,
determined experimentally, should be nearly constant for individuals
with valves of geometrically similar shapes and sizes yielding Re > 104.
Re is defined here by valve height and local free-stream velocity. The
ratio on the left side of (6.4) can be referred to as Arnold and Weihs's
ratio and should be useful in correlating slip velocities with size of
benthic animals in which Re > 104.
For live, unrestrained giant scallops (n = 17), &US = 11-37 cm s"1 of 2.7
to 12.0 cm valve height of animals was determined by Grant et al. (1993)
on a smooth acrylic floor. Independent data for the same species of live
scallop, but with the valves restrained with a rubber band at the anterior
end so that the band did not touch the smooth acrylic floor of the
Saunders and Hubbard recirculating flume were obtained by Wildish and
Saulnier (1993). The scallops (n = 26) were of 2.6- to 10.4-cm valve height
and gave a Us = 9-32 cm-s"1 (Wildish and Saulnier 1993), similar to the
results obtained by Grant et al. (1993). In addition to live giant scallops,
Wildish and Saulnier (1993) undertook experiments on model scallops
consisting of the valves of live specimens used in the previous experi-
ments, but with all wet tissue removed and the valves dried. Each set of
Behavioral Responses to Flow 245
o
30 -
Q?
CO
20-
Q
LU
LU -
K
Q_
CO
10-
CO oo/
0-
0 2 4 6 8 10 12
VALVE HEIGHT, cm
Figure 6.11 Valve height of the giant scallop as a function of slip speed: open
circles, live, shut valves, filled circles, dead, open valves (Wildish and Saulnier
1993).
hinged valves was then adjusted to a buoyant weight equal to that in life
by the addition of plasticine and the valves propped open to simulate the
observed maximal gaping in life - g in Fig. 6.10.
Results suggest that the slopes of Us as a function of valve height (Fig.
6.11) for live and model scallops are significantly different. Larger indi-
viduals (>8cm valve height) in which the valve is maximally gaped have
anonymously low slip speeds. This is linked to the increased frontal area,
which increases drag. In scallops of sizes >8 cm valve height, frictional
resistance with the bottom and buoyant weight is less efficient in over-
coming drag and lift. Possibly the behavioral phenomenon of recessing in
older scallops may be linked to this physical effect resulting from
allometric changes in the animal (Wildish and Saulnier 1993). Recessed
scallops are subject to lower drag forces because much of the body may
246 Benthic Suspension Feeders and Flow
this reference) because the adults link arms and passively filter feed by
extending their arms into the boundary layer. This has the effect of
anchoring them to resist energetic water movements and to create turbu-
lence in the benthic boundary layer and hence increase the feeding
potential.
An example of clumping behavior, which provides protection from
wave effects able to dislodge sessile epifauna, comes from a study of
sublittoral holothurians on the west coast of South Africa (Barkai 1991).
The holothurians involved were all passive suspension feeders. In loca-
tions where wave exposure was high, Thy one aurea (maximum density
377 nT2) and Pentacta doliolum (maximum density 464 m~2) occur to-
gether in dense aggregations. Thyone, which has degenerate tube feet,
insinuates its body beneath Pentacta, utilizing it as a means of attach-
ment. In more protected sites, Thyone creates separate clumps but
Pentacta is absent there. Flume studies showed that the strength of
attachment and resistance to increasing velocity as lift-off speed, Us - the
velocity at which a holothurian was swept away - were much higher in
P. doliolum than T aurea. Barkai (1991) considered that Thyone
commensally used the Pentacta colony to extend its range from protected
sites to more exposed ones. Model sea cucumber experiments in flumes
in which grouped and singular models were compared showed that Us of
grouped models was often two times higher than that of singular models.
The precise behavioral mechanisms involved in aggregation of these two
holothurians are unknown.
Usually, clumps are of individuals of a single species, e.g. polychaete or
amphipod tubes. Nowell and Church (1979) provide tube model experi-
ments in a flume that predicts tube density conditions which permit
reduction of drag within the clump. When tube density is sufficient to
give approximately one-twelfth the plan area of the bottom, the flow
changes to a "skimming flow" regime. In skimming flow over tube
clumps, much less kinetic energy reaches the bottom, and, under these
conditions, the flow-sediment interaction is stabilized. As pointed out by
Eckman et al. (1981), the Nowell and Church model does not accurately
predict stabilized versus de-stabilizing tube-sediment relations in field
conditions. They suggested that the discrepancy between field and labo-
ratory results was due to the common occurrence in the field of sediment
binding by mucus from bacteria or benthic diatoms. Eckman et al. (1981)
point out that a purely hydrodynamic cause of reduced kinetic energy
reaching the sediment as a result of tube clumping in field conditions has
not been demonstrated.
Table 6.9. Density and biomass among aggreging epifaunal suspension feeders.*
Bivalve molluscs
Modiolus modiolus Glacial till, Bay of Fundy, - 510b - 3,038b Peer et al. (1980)
Canada 281 1,023 Wildish and Peer (1983)
Abra alba Muddy sand, Consolidated 1,235 Warwick and Davies (1976)
substrate, Mud, Bristol 121
Channel, UK 70
Mesodesma deauratum Fine-medium sand, Grand 488 l,550 b 6,485 21,030b Hutcheson and Stewart
Banks, Nfld., Canada 3,010 12,000b 939 4,697b (1994)
Crassostrea virginica Estuaries of southeastern 4,077 214 Dame et al. (1984)
USA
Mytilus edulis Narragansett Bay, Rhode 2,139 10,962 Nixon et al. (1971)
Island, USA
Baltic Sea 36,000 158,000 101 Kautsky (1982)
Newfoundland, Canada 596 1,101 Thompson (1984)
Morecambe Bay, UK 1,234 Thompson (1984)
Perna perna South Africa 826 1,284 Thompson (1984)
Echinoderms
Amphiura filiformis Silt-sand burrower of 2,200 Warner (1979)
eastern Atlantic
Antedon bifida Outcropping rock, 1,200 Warner (1979)
Torbay, UK
Ophiothrix fragilis High tidal current flows 2,196 Warner (1979)
on various substrates
Polychaetes Fine sand-silt, La Jolla 500 15,000 Fager (1964)
Owenia fusiformis Bay, USA
Lanice conchilega Fine-medium sand, 20,200 1,094 Buhr and Winter (1977)
Weser estuary, FRG
Spio setosa Sand beach, Halifax 408 2,002 61 109 Muschenheim (1987a)
Harbour, Canada
Eudistylia vancouveri Rock, San Juan Is., 2,576 Merz (1984)
Pacific coast, USA
Phyllochaetopterus Fringing reef, Oahu, 62,400 Bailey-Brock (1979)
verrilli and Hawaii, USA
Mesochaetoptems
Sagittarius
Amphipods
Haploops fundiensis Silt-clay soft sediment, 376 923 0.35-0.53 0.96 Wildish (1984)
80-m depth, Bay of
Fundy, Canada
Ampelisca abdita Fine sand, Atlantic coast 1,360 73,000 Mills (1967)
Ampelisca vadorum Very coarse sand, 1,307 1,885 Mills (1967)
Atlantic coast,
N. Amer.
a
Wet biomass converted to dry by dividing by 3.33.
b
Maximum from single grab sample.
254 Benthic Suspension Feeders and Flow
Water level
0.5 second
Swimming
The pectinids are the bivalves with the best developed ability to swim,
although Stanley (1970) mentions solenids, solemyids, and cardids as
other bivalves with some locomotory abilities. Among scallops, two
types of locomotion can be recognized (Moore and Trueman 1971): a
predator escape reaction, in which a limited number of adductions of the
valves moves the animal as a result of the rapid expulsion of pallial water
which pushes the animal in the opposite direction from the point of
predator attack, and a complex swimming behavior.
Field studies have described complex swimming sequences in
Placopecten magellanicus (Caddy 1968) and Amusium pleuronectes
(Morton 1980) as (1) a rise from the bottom, (2) level flight, and (3)
passive sinking (Fig. 6.12). During the rise and level flight stages of
swimming, periodic contractions of the adductor muscle occur, depend-
ing on the length of the flight, the species, and possibly the size of the
individual. The timing of valve adduction is shown in the swimming
sequence of A. pleuronectes shown in Fig. 6.12. The force powering
swimming in scallops is provided by periodic jet propulsion. This in-
volves valve gaping to take in pallial water, closing the mantle edge
around the water, followed by rapid contraction of the adductor muscle,
which forces it out in two vectored jets dorsally on either side of the
auricles. Valve opening is usually described as being due to elasticity of
the ligament or resilium (Demont 1990), although Vogel (1985) makes a
good case for flow-induced forces augmenting the ligament in reopening
the valves. According to this argument, dynamic pressure on the right
valve ventral edge will be high and on the left valve low, thereby creating
Behavioral Responses to Flow 255
Re = - ^ (6.5)
Swimming
speed, cm s"1
h
Species cm xlO 4 F Mean Max Reference
(6.6)
w
The aspect ratio of the scallop valves is given by
(6.7)
(6.8)
where CL is the coefficient of lift for low aspect ratio wings obtained from
Hoerner (1975), 0.4 <CL< 0.8.
Complete data are available only for Placopecten magellanicus (Table
6.10). The data for this species show that the fineness ratio fits a near-
optimum shape, as does the relationship x/h = 0.6 in giant scallop adults
(where x = distance from ventral edge to the widest point in cross sec-
tion) for streamlining and fast swimming powers (Arnold and Weihs
1978). As pointed out in the section, Free-living epifauna, the blunt
258 Benthic Suspension Feeders and Flow
80
70
EED, cm-s" 1
60
50
40-
CO
O
2 30
^ 20
CO
10
20 40 60 80 100 120
L, mm
Figure 6.13 Swimming speed of Placopecten magellanicus as a function of valve
height: vertical lines, sample range; points, means, horizontal lines, one standard
deviation; curved line,fittedby eye (Dadswell and Weihs 1990).
trailing edge (dorsal side) opposed to the flow would be the optimum
hydrodynamic shape. During the ontogeny of the giant scallop, the ad-
ductor grows toward the mid-centre point of the valve, a process which
is complete by h = 5.5 cm (Dadswell and Weihs 1990). A similar ventral
displacement of the adductor muscle in relation to the valves is not
complete until h = 8.0 cm.
Swimming speeds shown in the table are approximate estimates made
from scatter plots of the original data made under non-standard environ-
mental conditions (e.g. temperature). Another difficulty in comparing
the swimming speeds shown in Table 6.10 is that they were determined
in different ways. Thus, Gruffyd (1976) and Morton (1980) measure the
time taken over a set distance during level flight swimming, while Joll
(1989) and Dadswell and Weihs (1990) use the total distance travelled
divided by the time taken from the initiation of stage (1) to the termina-
tion of stage (2) - that is the rise and level flight stages - but exclude the
Behavioral Responses to Flow 259
likely to swim when placed on clean sand than on seagrass beds, the
preferred habitat. Distances swum by bay scallops depended inversely
on the weight of fouling organisms, as simulated by adding 2-8 g of epoxy
putty to the left valve. The predacieous gastropods Fasciolaria tulipa and
Murex pomium always caused the bay scallop to swim when the tentacles
came into contact with the snail's foot. The direction of swimming by the
scallop was controlled by the location on the mantle edge contacted by
the predator - the escape reaction was directly away from the source of
stimulation (Winter and Hamilton 1980).
Summary
The foregoing shows that orientations to flow by a range of suspension
feeders, e.g. infaunal clams, phoronidians, brachiopods, barnacles, and
epifaunal bivalves, do take place. Research required to establish that
suspension feeders orient rheokinetically or rheotactically to enhance
their feeding includes observations of the behavior in field and lab and
experimental confirmation that the response is to flow and that the
individual derives trophic advantage from the preferred orientation.
Other species grow with respect to flow direction - rheotropisms -
thereby optimizing their feeding and growth. Research required to es-
tablish that rheotropic growth is occurring includes field observations
which establish a preferred orientation with respect to flow and flume
experiments which establish the effect of flow on growth. Although
relatively few suspension feeders have locomotory capabilities, those
that do may use them to increase their feeding opportunities. The re-
search required to establish this mechanism is often difficult to complete
satisfactorily. Steps involved in the process include field observations of
suspension feeder movement; experimental confirmation of sensory
cues, which frequently will not be flow-related ones; and some experi-
mental verification that restrained animals grow more slowly than unre-
strained ones.
Tube-living epifauna behavior has been little studied and so of the five
types of tube recognized here (there may well be others), studies have
only been started in one of the groups. The research required includes a
careful field observation of normal behavior in the tube, hydrodynamic
experiments to make clear the flow-tube interaction and the way ethol-
ogy is linked to it and that improved feeding or growth results from the
observed behavior. Abelson et al. (1993) used a physical model for
passive suspension feeders such as corals and sponges to predict the
Behavioral Responses to Flow 261
istics, consider that ontogenic changes are responsible for the size-
related swimming speed differences. In particular, the weight, which
scales according to h3 during ontogeny, outstrips the lift available, which
only increases as h2.
References
Abelson, A., T. Miloh, and Y. Loya. 1993. Flow patterns induced by substrata
and body morphologies of benthic organisms and their roles in
determining availability of food particles. Limnol. Oceanogr. 38: 1116
1124.
Anthony, K. R. N., and I. Svane. 1994. Effect of flow-habitat on body size and
reproductive patterns in the sea anemone, Metridium senile, in the
Gullmarsfjord, Sweden. Mar. Ecol. Prog. Ser. 113: 257-269.
Arnold, G. P., and D. Weihs. 1978. The hydrodynamics of rheotaxis in the
plaice (Pleuronectes platessa L.). J. Exp. Biol. 75: 147-169.
Ayling, A. M. 1976. The strategy of orientation in the barnacle Balanus
trigonus. Mar. Biol. 36: 335-342.
Bailey-Brock, J. H. 1979. Sediment trapping by chaetopterid polychaetes on a
Hawaiian fringing reef. J. Mar. Res. 37: 643-656.
Baird, R. H. 1958. On the swimming behaviour of scallops (Pecten maximus
L.). Proc. Malacological Soc. 33: 67-71.
Barkai, A. 1991. The effect of water movement on the distribution and
interaction of three holothurian species on the South African west coast.
J. Exp. Mar. Biol. Ecol. 153: 241-254.
Barnes, R. D. 1964. Tube building and feeding in the chaetopterid polychaete,
Spiochaetopterus oculatus. Biol. Bull. 127: 397-412.
Bertness, M. D., and E. Grosholz. 1985. Population dynamics of the ribbed
mussel, Geukensia demissa: the costs and benefits of an aggregated
distribution. Oecologia (Berlin) 67: 192-204.
Borowsky, B. 1983. Reproductive behaviour of three tube-building peracarid
crustaceans: the amphipods Jassa falcata and Amphithoe valida and the
tanaid Tanais cavolinii. Mar. Biol. 77: 257-263.
Bosence, D. W. J. 1979. The factors leading to aggregation and reef formation
in Serpula vermicularis L., p. 299-318. In G. Larwood and B. R. Rosen
(eds.) Biology and systematics of colonial organisms. Academic Press,
London.
Bowers, R. L. 1968. Observations on the orientation and feeding behaviour of
barnacles associated with lobsters. J. Exp. Mar. Biol. Ecol. 2: 105-112.
Buhr, K.-J. and J. E. Winter. 1977. Distribution and maintenance of a Lanice
conchilega association in the Weser estuary (FRG) with special inference
to the suspension-feeding behaviour of Lanice conchilega, p. 101-113. In
B. F. Keegan, P. 6'Ceidigh, and P. J. S. Boaden (eds.) Biology of benthic
organisms. Pergamon Press, Oxford.
Caddy, J. F. 1968. Underwater observations on scallop (Placopecten
magellanicus) behaviour and drag efficiency. J. Fish. Res. Board Can. 25:
2123-2141.
Carey, D. A. 1983. Particle resuspension in the benthic boundary layer by a
tube-building polychaete. Can. J. Fish. Aquat. Sci. 40 (Suppl.): 301-308.
264 Benthic Suspension Feeders and Flow
methods for the culture of giant scallops in the Bay of Fundy. Can. Tech.
Rep. Fish. Aquat. Sci. 1658.
Winter, M. A., and P. V. Hamilton. 1985. Factors influencing swimming in bay
scallops, Argopecten irradians (Lamarck, 1819). J. Exp. Mar. Biol. Ecol.
88: 227-242.
Witman, J. D., and T. H. Suchanek. 1984. Mussels in flow: drag and
dislodgement by epizoans. Mar. Ecol. Prog. Ser. 16: 259-268.
Young, G. A. 1985. Byssus-thread formation by the mussel Mytilus edulis:
effects of environmental factors. Mar. Ecol. Prog. Ser. 24: 261-271.
Yule, A. B., and D. J. Crisp. 1983. Adhesion of cypris larvae of the barnacle,
Balanus balanoides, to clean and athropodin treated surfaces. J. Mar.
Biol. Assoc. U.K. 63: 261-271.
Yule, A. B., and G. Walker. 1987. Adhesion in barnacles, p. 389-402. In A. J.
Southward (ed.) Barnacle biology. A. A. Balkema, Rotterdam.
7
Benthic populations and flow
270
Benthic Populations and Flow 271
SEDIMENT
PROCESSES
Bioturbation Burrown
Deposition
^ I. > / , "^Solutes
Physicar _ A I . v. A\
resuspensiqn ^otfbation
.;.;. .1 E r o s i o n - d e p o s i t i o n
cteria cycle
Fa
BENTHIC
W/////////////A Suspension
COMMUNITY Deposit Mixed Mixed Mixed Feeding
BENTHIC
PRODUCTION
Deposit --^"^ Suspension
Figure 7.1 Diagrammatic representation of benthic communities of the continental shelf along a gradient of
increasing water movement energy.
Benthic Populations and Flow 275
particles (4- to 62-jim median particle diameter) do not erode unless the
velocities reach nearly an order of magnitude greater than for medium
sand-sized particles. Depositional rates, on the other hand, are faster for
sand- than for clay-sized particles because the larger, more dense sand
particles settle at a greater rate than smaller, lighter clay particles. Be-
cause of biological activity involving binding of surficial sediments
(Madsen, Nilsson, and Sundback 1993), the figured critical sediment
erosion velocities in Fig. 7.2 are approximate.
NA = PRaC (7.1)
where
C is the seston concentration near the inhalant opening, or filtration
surface, g m3
P is the density of suspension feeders, number m"2
R is the average filtration rate of each individual in the population,
m^h"1
a is the average filtration efficiency (0% to 100%) of each individual
in the population
The term PRa is equivalent to the more familiar term clearance rate,
defined as cubic metres cleared per hour per square metre of reef. It has
the units of velocity, m h"1. The model shown in (7.1) does not deal with
seston deposition or resuspension. The flux is driven only by the suspen-
sion feeding of the benthos, which creates a concentration deficit near
the reef, and turbulent transport. Assumptions necessary are that the
suspension feeder reef is of semi-infinite extent; that the benthic bound-
ary layer is well developed, that is, that the velocity and concentration
profiles do not change in the downstream direction; and that there is no
mass transfer resistance attributable to a laminar sublayer.
278 Benthic Suspension Feeders and Flow
NA=yU(C0-C) (7.2)
where
V ;
C yU
^ = 1 + SDI (7.4)
SDI can be calculated for a particular site if the total clearance rate can
be estimated and typical velocities of the bulk flow are known. If "signifi-
cant" seston depletion is defined as C < 0.9C0, then it can be expected to
be observed if SDI > 0.11.
Data for sublittoral mussel reefs suggest that depletion effects can be
expected in the field. J0rgensen (1990) provides clearance rate data for
natural blue mussel reefs of the highest density found in the North Sea
and Canada as from 6-12m-h"1. Dare (1976) reported that a blue mussel
reef in Morecambe Bay, United Kingdom, of 1.4kg of soft tissue per m2
of reef had a total clearance rate of 12 m-h"1. Using this value for PRa
and y= 0.003, SDI can be calculated for a range of velocities (Table 7.3).
These results show that the highest density mussel reefs will deplete the
benthic boundary layer of seston at typical tidal velocities. Other
Benthic Populations and Flow 279
WMMmMMMM/W/MM/Mm
x=o
77/7T/
Figure 7.3 The concentration boundary layer which develops over a blue mussel reef in a unidirectional flow:
Z, water depth; X, mussel bed length; Co, initial seston concentration.
Benthic Populations and Flow 281
dc TT dc d ( dc | . , .
+ (7 - sinks + sources = 0 (7.5)
dt dx dzy dz)
where
c is the concentration of seston
U is the velocity at (x, z) and
Em is the vertical mixing coefficient
Equation (7.5), called the convective-diffusion equation, is of time aver-
aged form and does not consider changes in the cross-stream direction.
The first term represents temporal change in concentration of seston at
the point (x, z). The second term represents the change of advection of
seston by the average flow in the downstream direction. The third term
describes the vertical diffusion of seston due to turbulence. Several au-
thors, including Wildish and Kristmanson (1984), Smaal et al. (1986), and
Frechette, Butman, and Rockwell Geyer (1989), have proposed solu-
tions for the convective-diffusion equation.
Wildish and Kristmanson (1984) suggested an analysis of a developing
flume boundary layer in which the velocity boundary layer thickened in
the downstream direction so that velocity profiles could not be assumed
to be fully developed. The concentration boundary layer was assumed to
be well mixed in the vertical direction and to have the same thickness as
the velocity defect layer. The mean concentration in the boundary layer
was calculated by integrating the differential mass balance. This analysis
was extended by Verhagen (1986) and Smaal et al. (1986), who provided
an analytical solution and introduced a more realistic treatment of the
seston within the boundary layer. For the case of seawater >3 m deep
where steady unidirectional flows prevailed, using realistic profiles of
mean velocity and vertical eddy diffusivity, Verhagen (1986) also pre-
dicted the seston depletion to be expected at various downstream dis-
tances along the reef (Fig. 7.4). The abscissa can be considered as the
ratio of the velocity generated by the mussels (the number of mussels per
unit reef area times the volumetric flow per mussel) to the mean stream
velocity. It is therefore similar to the ratio SDI earlier proposed as a
measure of the tendency to depletion. An analytical solution is also
provided:
Co U (iDj
282 Benthic Suspension Feeders and Flow
0.6- x = 1000m
0.4-
O x = 500
\-
LLJ
a.
LJJ
Q 0.2-
= 333
1.0x10"3
where
x is the distance from the upstream boundary of the reef
C(x, 0) is the seston concentration at the inhalants
Dx is the vertical eddy diffusivity near the bottom
The units of the turbulent flux term are as in (7.1) and (7.2).
This solution is limited to small values of NRa5/D1 where 8 = seston
boundary layer thickness. Note that the depletion is directly propor-
tional to the product of two dimensionless groups, the abscissa of Fig. 7.4,
which depends on the population density and the current speed, and the
square root term, which is of the same form as the Peclet number defined
in Chapter 5 to express the ratio of advective transport to diffusional
transport. Depletion occurs when seston is moved rapidly from the
boundary by pumping, and smearing by turbulent diffusion cannot
smooth out the gradients thus created. The Peclet number term is deter-
mined by the mechanics of the flow and will become smaller if the
Benthic Populations and Flow 283
0.15 m-s"1. The effect of depletion is clearly shown. Time series mea-
surements at a sampling station near the centre of a mussel reef about
100 m long in the axis of the flow showed seston depletion at a depth of
0.05 m of approximately 50% at similar current speeds. Although the
model represents a simplification of the field conditions, the extent of the
agreement between the model and the fluorometric measurements is
impressive.
Bacterial
Bulk flow Temperature Mussel density ATP-seston nos. %
Expt. cm s"1 C number m2 % depletion15 depletion
a
Wet biomass 3.672kgm" 2 (Expts. 1-3), 4.848kg-m"2 (Expts. 4-6), 2.424kgm" 2 (Expts.
7-8).
b
ATP, adenosine triphosphate.
c
P<0.5.
Source: Wildish and Kristmanson (1984).
Table 7.5. Free-stream velocities and calculated SDI values for some
flume experiments with mussels.^
Biomass, u,
Species kg A F D W m " 2 cm s"1 SDI Reference
a
The following were used in the calculations: 7?=2.83L-h x for M. edulis and
R = 10.98 L-h" 1 for M. modiolus (Wildish and Kristmanson 1984, based on
J0rgensen 1990); y=0.006 (Butman et al. 1994); wet weight converted to
ash-free dry weight (AFDW) by multiplying by 0.1.
^ mmmi
10 0.05 0.5
^ T . 10 Qm o.3
CD 0 . 3 -
en
CD
Z3
E (
/
(/
c
oS /
CD
CE-0.1-
0 1 2 3 4
(7, cm s'1
Figure 7.7 Blue mussel growth rates (mean and range) as a function of velocity
(/, cm s"1) in 31-d-long concurrent experiments in Kirby-Smith growth tubes; 10
mussels per tube 364 mussels-m~2 with each mussel approximately 2.54 g wet
weight (Wildish and Kristmanson 1985); natural seawater from the Bay of
Fundy, Canada.
upstream (the first five mussels in the tube), mussels grew better than
downstream ones (Table 7.6) at flows <2cm-s~1, but at >2cm-s"\ there
was no significant difference between upstream and downstream indi-
vidual mussel growth. The theory predicts that not only velocity but also
seston concentration in relation to the composition of the suspension
feeder bed (mean size, density, and reef path length) affected average
growth, so this result is not universally valid. Thus, if the seston concen-
tration were increased or mussel density reduced, quite different growth
results could be expected.
To understand fully the ecological effects of flow on suspension
feeding populations, a wide flow range must be tested. Flow velocities
from -1-24 cm-s"1 were provided in a multiple-channel flume (Chap. 2)
with four separate channels in giant scallop growth experiments (Wildish
et al. 1987). Each channel had the same seston concentration but a
290 Benthic Suspension Feeders and Flow
a
Natural seston at a constant density of 10 mussels per tube = 364 mussels-m 2.
Source: Wildish and Kristmanson (1985).
0.4
o
o
0.3 -
0.2 -
0.1 -
0-
-0.10 i i i i i ih
0.24 - B
o 0.20 -
\
o
0.16 -
-fW 0.12 - o o*-^.
0.08 -
L
0.04- \
\
0- o
10 20
U, cm s'
Figure 7.8 Giant scallop Placopecten magellanicus specific daily growth rates of
adults (60 to 75 g wet weight) determined in 25- to 33-d-long experiments at
different times of the year, as a function of velocity (U, cm-s"1). Four channels
each with 16-24 scallops^ 18-27-m~2 were run concurrently in a multiple-
channel flume using natural seawater from the Bay of Fundy, Canada: A,
W = total wet weight, g; B, L = valve height, cm; N = number of days in experi-
ment (Wildish et al. 1987).
15-wk summer growth period, the effect of sediments (mud, sand, and
undisturbed control) at three different sites in the local area. Current
speed was measured at each site for such a small portion of the experi-
mental period that it is possible that the data are not representative of
conditions during the experiment. Grizzle and Morin (1989) concluded
that, whereas sediments had no effect on clam growth, site differences
did. Because environmental variables such as tidal velocity and seston
concentration were not controlled during the experiment, we consider
292 Benthic Suspension Feeders and Flow
that the positive relationship established between seston flux rate and
hard clam growth, based on three data points, may be fortuitous. Other
environmental variables such as seston quality, dissolved oxygen levels,
or inhibitory wind-wave effects could explain the higher growth rates at
one particular site. We concur with Grizzle and Lutz (1989) that there is
a need to determine the physiological responses of hard clams to velocity
and seston density in more controlled experiments.
Other authors (e.g. Cahalan, Sidall, and Luckenbach 1989; Wildish
and Saulnier 1993) have shown that feeding is not simply related to
sestonflux,but rather is interactively related to seston concentration and
velocity. In passive suspension feeders, too, Leonard (1989, and see
Chap. 4) found that in the crinoid he studied, the feeding responses were
to prey concentration and velocity separately.
In a 3-mo-long field experiment in the northern Gulf of Mexico, Judge,
Coen, and Heck (1992) tested the effect of seawater flow on hard clam
growth rates. The water depth was 0.6m and tidal range small (<15cm).
Wind-wave effects were an important driving force for flows at their
study locality. An open-ended four-channel flume, without a bottom,
was built in the field. The flume was 7 m long with 1.2-m-high walls, and
the flows within each channel were controlled by the width of the en-
trance. Comparative flows in each channel were assessed on one day so
that three treatments of 2% and 40% reduction, or 65% increase in
flows, were provided. Velocity was not measured continuously through-
out this experiment, although spot-check determinations suggested a
range of 0-27 cm-s"1. Six Mercenaria mercenaria were placed in the
Benthic Populations and Flow 293
middle of each channel in a plot of 0.25 m"2 (=24 clams m~2), but whether
other suspension feeders were present in the natural sand of the rest of
the flume was not clear. Since growth rates between all treatments and
the control were not significantly different, Judge et al. (1992) concluded
that velocity, independent of seston depletion, does not influence hard
clam growth. We believe that this conclusion is not justified by the
methods used, particularly because we know nothing of the temporal
distribution of flow throughout the experiment. This is important be-
cause of the stochastic nature of wind-wave effects and the demon-
strated ability of bivalves to adjust and compensate for the varying flow
periods (Wildish and Kristmanson 1988).
Comparative growth studies between siphonate (M. mercenaria) and
non-siphonate {Crassostrea virginica) bivalves were conducted by Griz-
zle, Langan, and Howell (1992) in a multiple-channel flume. Unfortu-
nately, only a limited range of free-stream velocities (<8cm-s~1) were
tested so the results must be regarded as preliminary. For 1991 results,
oyster growth (valve increments) was unimodally related to velocity with
a peak -lcm-s" 1 , while hard clam growth over the same flow range
showed only a and b of the unimodal response, with a broad b peak
between 2 and 4.5 cm-s"1.
Laboratory and field studies in Narragansett Bay, on the eastern coast
of the United States, were designed to determine the effect of flow,
seston concentration, and temperature on the feeding and population
biology of the acorn barnacle, Semibalanus balanoides (Sandford et al.
1994). In simple flow tank studies, the acorn barnacles were exposed to
unidirectionalflowsin the range 0-21 cm s"1 and at temperatures of 10-
25C, which were typical of local conditions. Barnacle feeding was in all
cases passive filtering, except at zeroflow,where slow cirral beating was
observed. The percentage of barnacles passively filtering was shown to
increase up to aflowof 21 cm s"1 and to peak at a temperature ~15C for
both adults and juveniles. Locally available seawater with its naturally
occurring seston resulted in variations in the percentage of barnacles
feeding so that the locality with the highest chlorophyll a values had most
barnacles feeding. These results were generally confirmed by growth
experiments in various local conditions. The study by Sandford et al.
(1994) might have benefitted from the use of an integrated environmen-
tal/physiological model for feeding and growth as suggested for the giant
scallop (Chap. 4, the section, Bivalves). The limited range of unidirec-
tional flows tested with respect to barnacle feeding and growth and
the absence of simulated wind-wave action, render laboratory results
294 Benthic Suspension Feeders and Flow
100-
50-
CO
Q 10-
DQ
5^
10 20 30 40 50
U av.
Figure 7.9 Deposit feeder () and suspension feeder (o) biomass (gwet
weight m"2) at seven estuarine or nearshore locations within the southwestern
Bay of Fundy, Canada (Wildish and Kristmanson 1979): x-axis, mean current
velocity as cm-s"1 over one tidal period at 1-2 m above the sediment-water
interface.
Location a b r N
a
x = g wet wt m"2 yr"1; y = mean velocity, cm s"1.
Source: Wildish et al. (1986).
Mean
Number growth per Total
mussels mussel growth
20 0.02 0.4
wwm/f,
10 0 0 5
/// //
rather than with total seston concentration. The disparity between this
and the previous results suggests a species-specific difference due to
resuspended sediment and its effect on bivalve growth.
0.7 A
el" 0.6-
CO
LO
Z3
E 0.5-
day1
0.4-
"S
0.3-
r
0.2-
oDC
CD
Maximum
Density range, Bed path growth
Species number m1 length, m decrease, % Reference
Andara granosa 125-2500 10.0 ?-70 Broom (1982)
Protothaca staminea 5-96 1.0 38-49 Peterson (1982)
Chione undatella 0.5-48 ]L.O 38-49
Circe lenticularis 5-30 ]L.O 40-90
Anomalocardia 5-30 1L.O -70 Peterson and Black
squamosa (1987)
Callista impar 10-60 ]L.O 10-30
Katelysia scalarina 20-160 ]L.O 35 Peterson and Black
(1988)
K. rhytiphora 20-160 1.0 40
Mercenaria 290-1159 0.6-1.2 50 Eldridge et al. (1979)
mercenaria
M. mercenaria 10-80 1.0 18 Peterson and Beal
(1989)
Eversole, and Whetstone 1979; Peterson and Beal 1989) and Australia
(Peterson and Black 1987, 1988).
Field experiments involving intertidal blue mussels Mytilus edulis have
shown that for patches of mussels, the individual "edge" mussels grew
faster than the control ones (Okamura 1986; Newell 1990). Here, the
oscillatingflowsof the shore enable "edge" mussels to be upstream for at
least some of the time, while the central mussels experience the seston
depletion effect and thus grow at slower rates.
XO=|YP^L (v.8)
and Risk 1979; Wildish 1980; Dobbs and Vozarik 1983; Shanks and
Wright 1986; Peterson et al. 1989; Nehls and Thiel 1993). Emerson (1989)
presented a statistical study of published secondary benthic production
data which included biotic (phytoplankton primary production, benthic
primary production, meiofaunal, macrofaunal, and total benthic second-
ary production) as well as physical variables (temperature, salinity,
depth, tidal height, sediment type, wind stress, and two wind shelter
indices). From 201 published data sets, a Pearson correlation matrix
determined that wind stress was weakly inversely correlated with total
(macrofauna, meiofauna, microflora) and macrofaunal secondary pro-
duction (r2 = 0.32 and 0.12, respectively). Benthic production was nega-
tively correlated with wind stress, suggesting that the overriding effect of
this variable in the benthic environment is negative, i.e. that increasing
wind stress at a locality results in decreased production. Further multiple
regression analysis on the log-transformed benthic secondary production
data (Emerson 1989) showed that wind stress, tidal height, wind shelter
indices, and seawater temperature explained -90% of the variance in
total secondary benthic production. This a posteriori analysis suggests
that water movement variables are major controls on benthic energy
flow, as is postulated in the trophic group mutual exclusion theory.
A spatial survey of the benthic macrofauna (biomass, numbers of
species) of high energy sediments on the coast of South Africa, inclusive
of intertidal and nearshore sublittoral regions, could be interpreted as
benthic limitation by water movement (McLachlan, Cockcroft, and
Malan 1984). In the wave breaking region of the high energy surf zone
transect, the extreme turbulence prevents any macrofauna from becom-
ing established, and it is only when more stable sediment-water inter-
faces occur that macrofaunal biomass and species numbers increase.
Emerson and Grant (1991) have observed one way in which wind-
wave activity could affect production on intertidal soft shell clam {Mya
arenaria) flats. Juvenile clams were eroded and transported across the
flats coincident with peaks of bedload transport associated with wind-
wave activity.
soft sediments of the Baltic Sea with two deposit feeding amphipods
{Pontoporeia affinis and P. femorata) and the facultative suspension
feeding bivalve Macoma balthica may be cited as an example. In the
Baltic, Macoma is absent or rare in those areas where dense populations
of Pontoporeia sp. occurred. Experimental tests in flowthrough aquaria
containing sediments (2-L glass jars) in which adult P. affinis and newly
settled spat of M. balthica were placed, showed that the latter decreased
with increasing densities of Pontoporeia. Apparently, P. affinis caused
direct physical injury to the spat, presumably ingesting the soft tissue,
although in nature the spat are not an important food source for
Pontoporeia.
Olafsson (1989) studied two contrasting localities in the Baltic contain-
ing Macoma balthica on the coast of southern Sweden. The first was in a
sheltered bay with limited water movement in a Zostera meadow where
the Macoma were deposit feeders. The second locality was exposed and
had a sandy sediment and more energetic water movement - mainly by
wind-wave action - where adult Macoma were primarily suspension
feeding. Adult density manipulation experiments in thefieldshowed that
juvenile Macoma growth and density levels were negatively affected at
the first, but not the second, location. Juvenile Macoma in both habitats
are deposit feeders, and thefieldresult was explained by dietary resource
overlap, which is absent where the water movements are more energetic
and where adult Macoma are suspension feeders. Experimental tests
suggested that Hydrobia sp., another deposit feeder, was negatively af-
fected by high densities of deposit feeding Macoma, whereas high densi-
ties of suspension feeding Macoma had no effect on deposit-feeding
oligochaetes of the second, water movement-energetic, sand habitat.
Andre and Rosenberg (1991) in manipulative field experiments in-
volving adjusting the densities of Cerastoderma edule and Mya arenaria
in the Baltic showed that adults reduced the settlement of bivalve larvae
and hence reduced the recruitment of new individuals proportional to
adult density. In a further laboratory flume study (Andre et al. 1993),
these findings were confirmed. Survival of settling larvae was propor-
tional to adult density of C. edule and increases in flume velocity caused
only a slight rise in predation risk.
There is as much evidence against the adult-larval interaction hypoth-
esis as there is for it (Posey 1990). Thus, deposit feeders, tube builders,
and suspension feeding clams do not always occur in discrete communi-
ties and may occur together in many natural situations (Posey 1990).
A direct field test involving manipulated densities of deposit- and
304 Benthic Suspension Feeders and Flow
Summary
The importance of benthic production limitation by flow, inclusive of
trophic group mutual exclusion, has been emphasized in recent reviews
(e.g. Olafsson et al. 1994; Hall 1994). However, flow influences food
supply to suspension feeders not only by controlling the turbulent supply
of seston, but by direct water movement impoverishment of both suspen-
sion and deposit feeders. The mechanisms involved have, by comparison,
received relatively little study and could include larval incompetency,
and adult washout during storm events.
Using the well-developed theory of hydrodynamics to help formulate
specific hypotheses about suspension feeders at all hierarchial levels has
proved to be, and we believe will continue to be, most rewarding. At the
level of the population the seston depletion effect (number 1, Table 7.2)
has been validated by both downstream reductions in seston concentra-
tion and downstream reductions in growth. Howflowand seston concen-
tration interact to produce increased production has begun to be docu-
mented for a few species (number 2, Table 7.2). The references cited for
number 2, Table 7.2, are observational confirmation of Hl5 although
experimental results (e.g. Wildish and Kristmanson 1985) which also
support it are available. More experimental results to cover a wider
range of variables (velocity, seston concentration) are needed. Another
ecological mechanism which may be important in the population re-
sponse to increases in flow involves the addition of suspension feeders,
by immigration and larval recruitment, and their subtraction, by emigra-
tion and death. All of these mechanisms allow three to four orders
of magnitude variation in suspension feeder production in the natural
marine environment, dependent largely on velocity and seston
concentration.
Of the four other hypotheses outlined in Table 7.2, there is field
evidence that the growth and production of suspension feeders can be
limited by energetic water movement (number 3), although the mecha-
nisms remain to be investigated in detail. The effect of suspension feeder
Benthic Populations and Flow 305
W 50 n
DC -3.0 Q
40-
O T
"D 8i
"2.0 Q w
ILTR,
CD 3 0 - DC =
|
u_ CO
-1.0
1
1 c:
o CO
20-
LU o
CO
CO I F
i O 10- LLJ
CO
^
LU
Z)
DQ
10 20 30 40
^inhalant
Figure 7.12 Blue mussel Mytilus edulis filtration rates, as percentage chloro-
phyll a consumed per hour. Dashed line (based on Wildish and Miyares 1990);
solid line, horse mussel Modiolus modiolus production (Wildish and Peer 1983).
The x axis shows ambient velocities at inhalant level for both species. Chla,
chlorophyll a.
the physiological feeding rate of one or a few individuals even though the
flow conditions used are the same. The change in hierarchical level from
the individual to the population is, of course, crucial in appreciating the
nature of this phenomenon. If feeding rates are an emergent property of
a population of suspension feeders, then methods presently used to
determine population feeding rates and hence growth, e.g. in bivalves
(Frechette et al. 1989; Frechette and Grant 1991), must be misleading.
This is because population level phenomena, such as growth, determined
in this way exclude recruitment mortality and the seston depletion effect,
and so cannot reliably predict population growth.
References
Aller, R. C, and R. E. Dodge. 1974. Animal-sediment relations in a tropical
lagoon, Discovery Bay, Jamaica. J. Mar. Res. 32: 209-232.
Anderson, F. E., and L. M. Meyer. 1986. The interaction of tidal currents on a
Benthic Populations and Flow 307
Background
The aquatic ecosystem concept of materials cycling, measured as energy,
carbon, or nutrient fluxes, was initiated by freshwater ecologists. Juday
313
314 Benthic Suspension Feeders and Flow
(1940) described the annual energy budget of an inland lake in which the
phytoplankton and attached aquatic plants utilized only about 1% of the
available subsurface solar energy during photosynthesis. Such energy
balances depend directly on the first law of thermodynamics - that
energy cannot be created or destroyed. Lindeman (1942), also working
in a lacustrine environment, showed how the food web complexity of an
ecosystem could usefully be simplified by categorizing the constituent
organisms into trophic-dynamic groups, hence, primary producers, her-
bivorous consumers, predators, and decomposers. Lindeman (1942) pro-
duced a diagram of a food cycle (web) for a lake in which the loss in
respiratory energy at each trophic level increased with distance from the
original solar energy supply. This conforms with the second law of ther-
modynamics because the transformation of energy at each trophic level,
e.g. light to chemical energy in plant tissue, involves an energy loss.
Both Juday and Lindeman tacitly assumed that solar radiation was of
most importance in freshwater ecosystems. Munk and Riley (1952) pre-
sented a quantitative theory that nutrient uptake by marine pelagic and
attached plants, e.g. seaweeds and seagrasses, was influenced by flow in
removing mass transfer limitation. Absorption of nutrients was consid-
ered to be dependent on the ambient concentration gradients and veloci-
ties. Thus, Munk and Riley (1952) showed how water movement, up to
a threshold level, was an integral part of materials cycling. The physi-
ological cast of their quantitative model, applicable to plant material
exchanges and water movement, precludes its central use in this chapter,
which is concerned with ecosystem level phenomena.
Odum (1971) and Odum, Finn, and Franz (1979) proposed that water
movement was unimodally related to attached plant biomass or produc-
tivity (Fig. 8.1) in an ecosystem level extension of part of Munk and
Riley's mass transfer limitation theory. Odum (1980) considered that
increases in attached marine plant production were caused by increases
in water movement which allowed the following:
Increased removal of waste gases and greater supply of bicarbonate for
photosynthesis
Increased nutrient fluxavailability to plant absorptive surfaces
Increased water column recycling of plant nutrients
Increased sediment-water interface fluxesleading to greater plant nutrient
cycling
The reference to this as a "tidal energy subsidy" by Odum (1980) is
clearly in error since no energy which subsidizes or enhances the produc-
tion process is directly supplied. Rather, the tidal energy supplied is
Ecosystems and Flow 315
Subsidy
LU
C/D
o
CL
CO
LU
DC Normal Range
LJLJ
Stress - \
CO
O Replacement
O \
LU
WATER MOVEMENT
Figure 8.1 Theoretical ecosystem response to water movement perturbation
(adapted from Odum et al. 1979).
Hydrodynamics
The theoretical account of Raupach and Thom (1981) concerning turbu-
lence in and above plant canopies provides a good starting point, al-
though these authors were dealing with terrestrial plants in air. Raupach
and Thom (1981) conclude that the subject of canopy turbulence is still
largely empirical. The plant structures which protrude into the boundary
layer, as well as the elasticity of the plants within the canopy, are of
importance in determining aerodynamic as well as hydrodynamic re-
sponses associated with plant canopies.
The fluid dynamics of seawater flow through a bushy red alga,
Gelidium nudifrons from Southern California, was studied by Anderson
and Charters (1982). Gelidium is found on semi-exposed shores, and
field measurements of orbital wave action of 2-10 cm-s 1 which passed
through the branches of the plant emerged as turbulent flow. Laboratory
studies in a low velocity water tunnel specially designed for this purpose
used dyes and a hot film anemometer to characterise flow. Gelidium
plants of 35 cm height and with branches of 0.05 cm thickness, when
tested with unidirectional flows in the water tunnel, damped turbulent
flows <6-12cm-s~1, producing a laminar exitflow.Atflows>6-12cm-s~1,
the flow leaving the plant was always turbulent.
A large kelp forest of 7- x 1-km area near San Diego, California, and
a nearby reference site which was kelp-free were compared by current
meter observations by Jackson and Winant (1983). The dominant kelp
was Macrocystis pyrifera, with some specimens reaching 40 m in stipe
length. These authors tested the null hypothesis that kelp plants had no
effect on flow patterns within the kelp forest (number 1, Table 8.1).
Vector measuring current meters were deployed at three depths in a
central location of the kelp bed and similarly at the reference site.
Records were obtained at both sites for a 7-d period and showed that the
average longshore currents within the kelp forest were -lcm-s" 1 . This
velocity was 43%-54% less than the root mean square velocity at the
reference site, thus supporting Hx in number 1, Table 8.1. Calculations
from these data suggest that seawater residence times within the kelp
Ecosystems and Flow 317
forest were >7 d. Measured velocities were always strongest at the kelp
forest edge, but slowed to average values within 100 m into the kelp
forest. In a further study of the same area, Jackson (1984) placed current
meters with temperature recorders along a transect across the kelp for-
est. As a result of the high drag of the kelp plants, present at densities of
0.02-0.14 plants-m 2 , internal waves which passed through the forest
were considerably dampened (Fig. 8.2), as evidenced by a smoothing of
high frequency fluctuations.
Understorey kelps dominated by Agarum fimbriatum were studied on
318 Benthic Suspension Feeders and Flow
'irfV^Uir-^^
+530 m
10 11 12 13 14 15 16 17
OCTOBER
Figure 8.2 Continuous temperature records over a four site deployment at
Point Loma kelp forest, near San Diego, California. The first site (-70 m) is 70 m
outside, the next (+120m) is 120m into the kelp, and so on (Jackson 1984).
the San Juan Archipelago, at 48N, 123E, by Eckman et al. (1989). The
relative degree of water movement activity, assessed by the gypsum flux
method, showed that the kelp plants impeded flow and dampened vari-
ability in currents relative to that of a reference location nearby. As with
the California data outlined previously, the effect was accompanied by a
reduction in mass transport and shear within the kelp bed. Because of
the reduced flow within the kelp, Eckman et al. (1989) reasoned that
increased sedimentation should occur (Table 8.1). Sedimentation rates
were measured by collection of the particles on acrylic plastic (Plexiglas)
plates, followed by pumping into deflated plastic bags,filtering,and then
weighing. Because this method did not properly distinguish sedimenta-
tion and resuspension events, a particle impaction experiment was also
employed. It involved glass beads of 22 (im diameter and with a settling
velocity of 0.0276 cm s"1, which were slowly released upstream of micro-
scope slides coated with vacuum grease and on which the beads became
trapped. Results suggested to Eckman et al. (1989) that particle trapping
was reduced in the presence of kelp plants, consistent with a reduction in
the intensity of turbulent mixing within the canopy. The increased sedi-
mentation within kelp canopies is explained by the greater retention
times that particles must spend within the canopy and hence a greater
proportion of particles which must settle. During flood-ebb tidal
Ecosystems and Flow 319
Materials cycling
A functional account of the Laminaria-dominated kelp forests of Atlan-
tic Canada is given by Mann (1973). From this and related studies, a
partial carbon budget can be constructed (Fig. 8.3). Such a budget repre-
sents the following:
An estimate of the amount of energy passing through the ecosystem
A simplification of a more complex cycle than is shown in Fig. 8.3
Hypotheses about the pathways of energy, carbon, or minerals in the kelp
ecosystem
Salient features of this budget are the very high primary production
levels achieved by Laminaria longicornis, L. digitata, and Agarum
cribosum, which are based on energetic water movement, principally
wind-wave action, which enhances photosynthesis by supplying and re-
moving gases, bicarbonate, and nutrients. The proportion of dissolved
organic matter (DOM) utilized in the pelagic microbial loop was not
estimated, although it was probably small. Both DOM and particulate
organic matter (POM) are exported from the kelp forest. Estimates of
net export of POM suggest that it is a large proportion of the total
because the principal herbivores, the sea urchin Strongylocentrotus
droebachiensis and periwinkle Littorina littorea, utilized <10% of the
available kelp carbon. High density aggregations of urchins sometimes
occurred and could be very destructive of the kelp plants by biting
through the stipe bases, thus forming cleared patches referred to as "sea
urchin barrens" (Mann 1973). Whether the feeding aggregations of S.
droebachiensis are controlled by decapod key predators, as suggested by
Mann (1973), or by some other mechanism, inclusive of innate urchin
behavior (Hagen and Mann 1994), is still a matter of debate.
In field studies conducted on the Aleutian Archipelago, Alaska, Estes
and Palmisano (1974) chose two study sites: Rat Island at 52N, 178E,
and Near Island at 52N, 174E. The two sites were 400km apart but
320 Benthic Suspension Feeders and Flow
[930-1150] TURBULENTTRANSPORT
marked difference between the kelp-free site at Near and kelp-rich site
at Rat Island was due to the presence of a key predator, the sea otter
Enhydra lutris, at the latter (Table 8.1, number 3). High densities - 20-
30 sea otters km"2 - removed the larger sea urchins, thereby preventing
destructive damage by them to the kelp. Circumstantial field evidence
gathered in support of this hypothesis was limited; it included the follow-
ing: sea urchin densities were much higher at the kelp-free sites; at those
sites where sea otters were present at high density, only small urchins,
less than 35-mm test diameter, were found; and some differences in
higher trophic level species between sites, e.g. the presence of harbour
seals and bald eagles, were related to the higher primary production
available at kelp-rich Rat Island.
Further study at these sites was undertaken by Duggins, Simenstad,
and Estes (1989) to determine the importance of detrital pathways
within coastal ecosystems (Table 8.1, number 4). Comparative field
measurements of growth for two suspension feeders, Balanus glandula
attached to test plates and Mytilus edulis, were made by locating marked
individuals sublittorally within kelp beds at Rat Island and at the same
depth in the kelp-free environment at Near Island. After 1 yr of growth,
the rates were compared (Fig. 8.4); barnacle growth was five times and
mussel growth two to four times greater at the kelp-rich site. Because the
carbon ingested during feeding by these suspension feeders could have
originated from sources other than kelp, e.g. from phytoplankton, it was
necessary to use a method capable of tracing the feeding habits of sus-
pension feeders. Because the 8 12C: 13C signatures of phytoplankton and
kelp were so different, it was possible to indicate the source of the carbon
on which they had been feeding. Of 11 species examined by this tech-
nique from the Alaskan sites, Duggins et al. (1989) found that 10 species
had enriched levels of 8 13C from Rat Island, suggesting that they had
been feeding indirectly on kelp (on average 58% C from kelp). Only
Mytilus edulis did not show enrichment in this way, suggesting that it may
prefer a phytoplankton diet. By contrast, 8 13C enrichment at the Near
Island site was much lower (on average 32% C from kelp).
Both field growth experiments and measurements of carbon isotope
ratios in naturally co-occurring benthic macrofauna of kelp beds led
Duggins et al. (1989) to conclude that the kelp plants themselves are a
significant carbon source for them. Because of the nature of the measure-
ments, the results give no indication about how important this pathway is
to suspension feeders. Nor do they allow an estimation of the relative
importance of kelp carbon export from the kelp forest.
322 Benthic Suspension Feeders and Flow
43
KELP-RICH KELP-FREE J
Figure 8.4 Growth experiments on the Aleutian archipelago, Alaska: A,
Mytilus edulis; B, Balanus glandula (from Duggins et al. 1989). Two kelp-rich
sites from Rat Island, one kelp-free site from Near Island. Number above each
bar, sample size.
local water movements are included shows that bacteria have insufficient
time to accumulate before being advected from the kelp forest.
Kelp productivity
Overstorey kelps are present on rocky substrates of the nearshore
coastal environment to a maximum depth of -40 m. Hence, they form a
narrow zone close to the shore and are light-limited at greater depths.
Primary production estimates made here suggest very high levels, e.g.
1750 g C-m^-yr"1, for a Laminaria-dominated forest on the Atlantic
coast of Canada (Mann 1973). A West Coast kelp forest dominated by
the sea palm Postelsia palmaeformis is reported (Leigh et al. 1987) to
have an even higher productivity, 3000gC-m"2-yr~1, on the most wave-
exposed parts of the coast of Washington.
A cursory review of the data concerning kelp productivity in relation
to wind-wave exposure suggests apparently conflicting results. Thus,
Laminaria kelp forests were less productive at the most exposed sites on
the Atlantic coast of Canada studied by Gerard and Mann (1979), and
Macrocystis sp. beds grew best at intermediate levels of wave exposure
on the southern coast of Chile (Dayton 1985). Yet, on the coast of
Washington, Postelsia sp., referred to earlier, as well as the shrubby
kelps, Lessoniopsis littoralis, are found only on the most wave-exposed
locations (Leigh et al. 1987). Unfortunately, although Leigh et al. (1987)
determined kelp productivity at a number of different sites on the Wash-
ington coast, the ranges of productivity were lumped and not related
individually to an estimate of wave power, thus rendering the results
unconvincing.
An obvious difficulty in comparing the work of different authors on
kelp productivity is the lack of a commonly used absolute measure of
wave exposure. This problem was considered in Chapter 2 and the
gypsum dissolution method proposed is an inexpensive way of compara-
tively measuring wind-wave activity.
Despite the lack of a wave exposure measure for the data cited, we
believe that the gross differences in exposure suggested are real and
result in the responses by kelp forests suggested. If this is so, the key to
understanding the conflicting results lies in the probability that kelps are
adapted to different amounts of wave exposure, both inter- and
intraspecifically. Indeed, the form of a sheltered versus an exposed frond
of Laminaria longicornis (Fig. 8.5) bears this out. Further examples of
intraspecific lamina form changes, e.g. within Macrocystis sp., which are
Ecosystems and Flow 325
Figure 8.5 Growth forms of Laminaria longicornis from the Canadian Atlantic
coast: A, sheltered site; B, exposed site (Gerard and Mann 1979).
For the other hypotheses shown in Table 8.2 (numbers 2-5), there is
some evidence for the alternative hypotheses, as described in the listed
references.
Hydrodynamics
Laboratory studies in a recirculating flume 6 m long by 0.15 m wide by
Fonesca et al. (1982) were designed to determine the effect of Zostera
marina shoot density on flows within the meadow. This work was criti-
cized by Gambi, No well, and Jumars (1990) because the flume did not
simulate dynamically similar conditions to those present in the field, in
particular, the flume was too narrow, resulting in flow blockages when
seagrass shoots were present.
A larger annular flume with an 8-m-long working section, 2 m wide,
was used by Gambi et al. (1990). This flume had a water depth of 25 cm
and a 4-cm-deep sand layer on the flume floor into which young Z.
marina plants were inserted. The eelgrass plants used were -16 cm in
height so that flow could form above the bed. The experimental eelgrass
patch (15 by 100 cm long) was placed where the flume boundary layer
was fully developed and so that the plants were 30 cm from the walls.
Flow was deflected around and over the patch so that there was a sharp
interface near the shoot tips where high shear and turbulence intensity
occurred. Within the eelgrass shoots, velocity was reduced as a function
of shoot density and distance into the meadow (Table 8.3). In these
experiments, the penetration offloweffects into the eelgrass bed is more
marked than the density effect.
Field studies by Ackerman (1983), Fonesca et al. (1983), and Eckman
(1987) also suggest that reduced velocities occur within natural Zostera
meadows. Calculations by Eckman (1987) to determine the flux rates
through a high density (604-1134 plants m~2) and a low density (183-271
plants m"2) eelgrass meadow showed that at slower free-stream veloci-
328 Benthic Suspension Feeders and Flow
Freestream velocity 5 cm s
1200 20.7 38.3 43.1 35.2
1000 11.0 31.4 26.7 40.6
800 7.55 19.9 21.4 28.9
600 3.58 12.3 13.7 20.3
400 9.6 18.7 29.7 21.8
Freestream velocity 20 cm
1200 8.9 16.6 19.9 34.2
1000 10.4 19.6 23.16 21.0
800 2.2 13.4 21.0 18.3
600 5.2 13.9 21.7 19.8
400 2.0 8.3 15.2 14.7
ties, there was a 16% reduction in flux due to the higher density. At a
higher velocity, difference was increased to 52%. The field results do not
agree very well with the flume results shown in Table 8.3, perhaps be-
cause the latter are artifacts of the small bed size tested.
Field observations with a dual axis electromagnetic current meter
suggested to Ackerman and Okubo (1993) that the plants themselves
were generating mechanical turbulence. The frequency generated by the
eelgrass plants of 6.4-8.0 s"1 results because they are hydroelastically
translating fluid energy.
The ecosystem implications of seawaterflowsto and in seagrass mead-
ows include larval supply (Chap. 3), effects on seagrass production,
nutrient exchange between sediment and seawater inclusive of re-
suspension and deposition, direct shear stress on seagrass plant stability,
Ecosystems and Flow 329
Seagrass productivity
The effect of tidal flows on seagrass productivity (Table 8.4, number 2)
was investigated, apparently independently, by Conover (1968) on
eelgrass meadows of the U.S. East Coast and by Odum (1961,1971) and
Steever, Warren, and Niering (1976) on cordgrass meadows on the same
coast.
In Conover's study, eelgrass standing crops of various locations are
presented as a function of the mean hourly tidal velocity averaged over
a complete springs to neaps period (Conover 1968). Where the current
meter was deployed in this study was not stated; presumably free-stream
velocities seaward of the eelgrass meadow are meant. The data for
Charlestown Pond, Rhode Island, are shown in Fig. 8.6. Standing crop
data in Fig. 8.6 are probably a good indication of eelgrass primary pro-
ductivity because herbivores are unimportant in this ecosystem. A char-
acteristic unimodal response to velocity is seen. Conover (1968) sug-
gested that the positive relationship could be explained by an increased
photosynthetic activity as flows increased, and that the inverse relation-
ship was due to an inhibition of metabolism at high flows and/or the
shearing effect of high flow, which prevented the plants from becoming
established. Further results from Texas and from Massachusetts and
Rhode Island along the U.S. East Coast for both seagrasses and
seaweeds (Fig. 8.7) show a similar trend. The data confirm that individual
species of canopy-forming macroalgae have distinctively different re-
sponses to tidal flows, confirming the suggestion made previously for
kelp.
Steever et al. (1976) compared the standing crops of Spartina
alterniflora at various marshes along the Connecticut, United States,
coast. For 10 such marshes in 1971, the standing crop dry weight of
cordgrass was linearly related to mean tidal height (how this was esti-
mated was not stated) with a calculated slope equal to 544 g dry
330 Benthic Suspension Feeders and Flow
CL 200 -
O
D C CM
U'E
CD
\
W
,*
D)100 - \
CO
25 50 75
U, cm s"
Figure 8.6 Standing crop of Zostera marina in Charlestown Pond, Rhode Is-
land, 1962-63, as a function of velocity (average lunar period velocity expressed,
cm-s" 1 ) at different sites (Conover 1968).
200
Figure 8.7 Standing crop data for seaweeds and seagrasses for various locations
on the U.S. East Coast where summer temperatures are ~23-28C and salinities
are -29-31 o/oo; dashed line divides lagoon (to left) from channel and coastal
ecosystems: A, Ruppia maritima; B, Thalassia testudinum; C, Stylophora
rhizoides; D, Zostera marina; E, Chondrus crispus; F, Laminaria agardhii; G,
Bryopsis plumosa (Conover 1968).
332 Benthic Suspension Feeders and Flow
flux into or out of the seagrass meadow is complex, but generally is a net
importer during the summer when the Spartina plants are growing and a
net exporter during winter. Wolaver et al. (1983), working on another S.
alterniflora meadow in Virginia, United States, also found pronounced
seasonal differences in fluxes to and from this ecosystem. These studies
were aided by deploying a field flume (see Chap. 2) to aid in sampling
tidal fluxes to and from the seagrass meadow.
The extensive body of work on nutrient fluxes to and from seagrass
meadows has been reviewed by Hemminga, Harrison, and van Leut
(1991). The inputs include nitrogen fixation, sedimentation, and direct
uptake from seawater advected to the marsh. The losses include diffu-
sion and tidal transport away from the meadow, denitrification, export of
POM, loss due to herbivory, and possible transport away from the
meadow, as well as direct exudation of nutrients from seagrass roots and
leaves.
Seagrass stability
Qualitative observations regarding the stability of seagrass meadows
have been made by Scoffin (1970), Patriquin (1975), Orth (1977),
Fonesca and Kenworthy (1987), and Fonesca et al. (1983). In general, the
"blowouts" - crescent-shaped areas of erosion (Fig. 8.9) where water
movement has torn away the seagrass plants - are associated with iso-
lated patches or mounds of seagrass on the seaward face of the meadow.
Ecosystems and Flow 337
As classified by Fonesca et al. (1983), these areas are also associated with
the highest water movement energies.
The blowouts are usually associated with extreme wave energy, as may
occur during storms and may result in uprooting and formation of large
drifting mats of seagrasses in the Gulf of Mexico and the Caribbean
(Patriquin 1975). Scoffin (1970) described blowouts in Bimini Lagoon
which were caused by locally high tidal velocities. Preliminary observa-
tions made by Scoffin (1970) in an underwater flume (Chap. 2) suggest
that the shear forces associated with unidirectionalflowsresult in erosion
of sediment around the roots of Thalassia testudinum in the free-stream
velocity range of 50-150 cm-s"1.
Tidal flow
(cm- Bivalve Bivalve Bivalve growth rate
density path length
Seagrass meadow Bivalve ^max (no. m"2) (m) Seagrass Reference References
"Eelgrass" Mercenaria mercenaria >20 8 69-101 0.6-1.2 Low High Kerswill (1949)
Zostera marina and M. mercenaria 7 0.7-1.3 9 High Low Peterson et al.
Halodule wrightii (1984)
Z. marina and M. mercenaria 27 0.3 49 1.0 High Zero Irlandi and
H. wrightii Peterson (1991)
Z. marina Argopecten irradians 13-98 ? 22 Low High Eckman (1987)
Z. marina Anomia simplex 13-98 ? 35 Same Same Eckman (1987)
Ecosystems and Flow 339
outside a seagrass meadow (Table 8.5), include the food subsidy hypoth-
esis of Table 8.4 (number 5) within seagrass meadows and the physiologi-
cal cost for bivalve feeders to maintain their position on the shore, which
is proportional to local velocities (number 6). Field experiments under-
taken by Lin (1989) in a North Carolina Spartina alterniflora meadow
involved the ribbed mussel Geukensia demissa, which occurred in a 1- to
2-m wide strip on the seaward side. Manipulative experiments involving
singleton or small groups of 10 or 30 mussels were undertaken at differ-
ent tidal levels and positions within the seagrass meadow. Seston deple-
tion effects could be excluded as an explanation for differences in growth
because treatments at the same tidal level gave identical results. At
one of the two sites studied, ribbed mussels grew much faster in the
middle of the seagrass meadow than at the edge, consistent with pres-
ence of a food subsidy (number 5, Table 8.4). Judge, Coen, and Heck
(1993) found that seston concentrations were enriched at <5cm above
the sediment-water interface. The chlorophyll-containing seston in these
samples was made up of 90% pennate diatoms, although an estimate of
non-living POM was not made. For both of these studies the link show-
ing that the living or non-living POM enhanced bivalve growth was not
made.
That the energetic cost of maintaining shore position is higher where
water movement energies are higher and sediments less stable (Table
8.4, number 6) has been suggested repeatedly. Irlandi and Peterson
(1991), for example, were able to show by seagrass-canopy trimming
experiments that this reduced hard clam growth, consistent with in-
creased velocities in these plots and circumstantial evidence to support
Hi in number 6. They also observed that hard clams grew significantly
more slowly at the seagrass meadow edge where higher velocities were
also present.
Because multiple limiting factors are probably involved in interpreting
the Table 8.5 results, one should also consider biotic factors. The effect
of seagrass beds on predator-prey relations was reviewed by Orth, Heck,
and van MontFrans (1984). Effects could be separated into those involv-
ing the seagrass root rhizome and others involving the shoots. Peterson
(1982) showed that seagrass meadows provide sufficient below-ground
structural complexity to hinder the efficiency of whelk predators,
Busy con sp., capturing Mercenaria mercenaria. Siphon nipping of the
suspension-feeding bivalve Protothaca staminea by predaceous fish was
found to be 9-11 times greater outside the seagrass meadow than within
it (Peterson and Quammen 1982). Simulated siphon nipping of M.
340 Benthic Suspension Feeders and Flow
Bivalve reefs
The potential for suspension feeding bivalves to form dense
aggregations, or reefs, occurs in suitable locations throughout the world.
The reefs are most common at intertidal or nearshore subtidal locations,
particularly in estuaries or other areas of high productivity or tidal en-
ergy. Common bivalves which form reefs include the mytilid mussels and
oysters (Ostreidae).
Hydrodynamics
We know of no studies that have set out to determine specifically the
hydrodynamic variables within a natural bivalve reef, although some
have done this incidentally in determining filtration rates of the bivalves.
Such studies include those of Frechette and Bourget (1985) and
Frechette, Butman, and Rockwell Geyer (1989) in the field, Butman et
al. (1994) in a laboratory flume, and Monosmith et al. (1990) for model
bivalve populations in a flume.
The study by Frechette et al. (1989) concerned an intertidal mussel reef
which was 30 m wide by 100 m long in the St. Lawrence estuary, where the
tidal range was -4.8 m. The average biomass for the Mytilus edulis reef
was 500 g dry-m"2. Seawater sampling to determine phytoplankton con-
centration and measure filtration rates by standard physiological meth-
ods was at three heights above the reef: 1.0,0.5, and 0.02-0.05 m. Samples
were drawn continuously by a centrifugal pump at 3.5L-min~1 in a 1.3-
cm-inner-diameter (ID) plastic hose. The current velocities were deter-
mined by continuously recording Aanderaa current meters fitted with
salinity and temperature probes. Current meter rotors measured veloci-
ties at 1.0 and 0.01-0.1 m above the reef. By using benthic boundary layer
theory (Schlichting 1968) and velocity data, the logarithmic velocity
profile was constructed; from it the friction velocity, /*, and bottom
roughness parameter, Zo, were determined. From their detailed hydrody-
namic model analysis, Frechette et al. (1989) were able to show that the
commonly observed tendency for the mussels to form hummocks by
overgrowing other mussels may result in bottom roughness parameter
increases sufficient to increase turbulent transfer of seston to the reef
significantly.
342 Benthic Suspension Feeders and Flow
Reef metabolism
Studies which have measured bivalve reef metabolism (Table 8.7) are
considered here in detail. The aim is to determine whether bivalve reef
oxygen uptake data support the conclusions of Chapters 4 and 7 that
increased seawater velocity can enhance bivalve growth by increasing
the turbulent supply of sestonic food. The results summarized in Table
8.7 are of total metabolism of the reef, inclusive of photosynthetic oxy-
gen production, e.g. from phytoplankton in seawater or diatoms on the
reef surface. The only seasonal study of metabolism of which we are
aware is presented by Dame et al. (1992) for a South Carolina oyster
reef. This study found net oxygen production in winter when macroalgae
were present and denoted by a minus sign in Table 8.7; during the
summer, the oyster reef was strongly heterotrophic with a net oxygen
demand (BOD). Physiological data for oyster oxygen uptake rates dis-
cussed in Dame, Spurrier, and Zingmark (1992) suggest that the oysters
are only responsible for 10% of the annual reef oxygen uptake. Other
uptake sources are aerobic bacteria, meiofauna, other macrofauna, and
the chemical oxygen demand (COD) of the reef sediments.
During summer observations of a Narragansett Bay mussel reef,
Nixon et al. (1971) attempted to determine the effect of tidal velocity on
reef metabolism. Dissolved oxygen determinations in triplicate were
measured by Winckler titration. Seawater samples were obtained up-
stream of the seaward edge of the reef and downstream in the dominant
tidal direction, some 20-25 m away. Depth at which the samples were
taken was not stated, and a crude measure of velocity was made by
observing dye or drift bottle movements, so that velocities may be inac-
curate. The data in Fig. 8.10 were obtained by subtracting the down-
stream from the upstream dissolved oxygen onfloodingtides. The differ-
ence - assuming a well mixed tidal prism above the mussel reef -
represents the oxygen uptake calculated for the area over which the
water mass has moved in its passage between the two sampling points.
The instantaneous rate of BOD, /, must be corrected for the volumetric
flow rate which passes over the reef (Chap. 2, the section, Field flows).
Do the results of Fig. 8.10 support the conclusions of Chapters 4 and 7
and the thesis that water movement energy is a community metabolism
multiplier (Table 8.8, number 1)? Assuming that the reef filtration rates
are directly proportional to metabolic uptake of oxygen, the results do
support the theoretical curve of Fig. 4.1 for a bivalve with inhalant
normal to theflowand also the empirical fits of Figs. 7.7 and 7.8 based on
Table 8.7. Comparison of independent studies to measure bivalve reef metabolism.
Variable Nixon et al. (1971) Dankers et al. (1989) Dame et al. (1992)
Location Narragansett Bay, Rhode Island of Texel, North Inlet, South Carolina,
Island, U.S.A. Netherlands U.S.A.
Dominant species Mytilus edulis M. edulis Crassostrea virginica
Temperature (C) 20.1-23.5 14-16 8-30
Salinity (o/oo) 27-30 ? 15-35
Reef area (m2) 100 ? 600
Path length of reef measured (m) 20-25 10 10
Bivalve density (m~2) 6261 7 7
Bivalve dry meat biomass (g m~2) 1523 818 196
Reef respiration rate (g O2 m"2 h"1) 0.2-2.8 1.2-3.9 -1.5 to +4.0
344 Benthic Suspension Feeders and Flow
8-i
v 6-
CM
' 4-
C\J
o
oO)
^r 2 -
o
>
0-
10 20 30 40 50 60 70
1
I/, cm s"
Figure 8.10 Instantaneous rate of oxygen uptake (/) of a Narragansett Bay blue
mussel reef as a function of free-stream velocity, U, cm-s"1: solid dots, field
derived data; open dots, recalculated laboratory data. The fitted hyperbola has
r = 0.83, with single arrowed point excluded (from Nixon et al. 1971).
c/)
oysters fed on the flood tide but not on the ebb, when much of the seston
had already been removed by the oysters'filtering.This interpretation by
Dame et al. (1992) is consistent with other field studies which linked tidal
cycles with feeding as indicated by digestive cycles. Thus, mussels M.
edulis (Langton 1977) and sublittoral Pecten maximus (Mathers 1976)
fed on the flood tide. In the latter case, two distinct subpopulations of
scallops were present - one facing into and feeding on the flood, the
other away from the flooding current and feeding on the ebb current.
Histological studies by Mathers (1976) showed that the digestive physiol-
ogy of each subpopulation was ~6h out of phase, supporting a tidal
periodicity of digestion. Langton and Gabbott (1974) also described a
346 Benthic Suspension Feeders and Flow
tidal periodicity of feeding in oysters Ostrea edulis grown near low tide
on the shore at Menai Bridge, North Wales, by monitoring the crystalline
style physiology.
In the study described by Dankers, Dame, and Kersting (1989), two
methods of measurement were compared: a 200-m2 concrete tank sup-
plied with mussels and natural seawater and part of a natural blue mussel
reef enclosed by the tunnel method of Dame, Zingmark, and Haskin
(1984). Results of the former gave average net uptakes of 0.5-1.0g
O2 m~2 h"1, which were somewhat lower than similar values for a natural
mussel reef (Table 8.7). The discrepancy is explained by the sensitivity of
the results to the timing of measurements. Hence, net oxygenfluxeswere
low during the day when photosynthetic production of oxygen was oc-
curring. Nighttime observations were correspondingly high because this
production was absent.
An independent estimation of mussel oxygen uptake was made by
Dankers et al. (1989) from standard physiological data linked to biomass.
This yields a value of 0.37mgO2-g~1*h~1 equivalent to an uptake of
0.3 g O2 m~2 h"1 by the mussel reef. This represents only 8% to 25% of
the day-night dissolved oxygen flux, as determined by the tunnel
method.
Pelagic-benthic coupling
A direct link between the surface primary productivity of the oceans and
benthic productivity, or biomass, of the same area was proposed by
Rowe (1971). From published data he showed that surface primary pro-
ductivity ranked second only to depth in controlling benthic biomass.
Thus, carbon fixed photosynthetically in surface waters, and after pas-
sage through a pelagic food web, reached benthos in the form of faecal
pellets (Hargrave 1973). Rowe et al. (1975) also showed that the ex-
change was a two-way process with benthically regenerated plant nutri-
ents, e.g. various forms of nitrogen, recycled to primary producers in
shallow surface coastal waters. It was thus possible to speak of pelagic-
benthic coupling involving positive-negative feedback loops between
primary and secondary benthic producers.
Typically, energy or carbon budgets are calculated on the basis of unit
area - usually 1 m2 - close to the area actually sampled by a randomly
placed grab or corer used in estimating soft sediment benthic biomass
and hence secondary benthic production. Primary production, too, is
Ecosystems and Flow 347
estimated on the basis of unit area or volume. Thus, the first efforts to
link pelagic and benthicfluxeswere made on the basis of small vertically
linked areas. A simple regression model proposed by Hargrave (1973)
was successful in predicting the magnitude of the annual sediment oxy-
gen demand on the basis of the amount of pelagic primary production
and the mixed layer depth of the water column at the same site.
Yet, some authors found discrepancies in balancing energy budgets if
a tight vertical link between pelagos and benthos was assumed. Thus,
Bernard (1974), who described a partial energy budget for an oyster
Crassostrea gigas reef in British Columbia, showed that the energy avail-
able from the local primary production was much less than the oysters'
allocation of energy to the annual production of gametes. Emerson,
Roff, and Wildish (1986) provided a partial carbon budget based on one
site in the Bay of Fundy (Fig. 8.12) - a macrotidal estuarine ecosystem in
eastern Canada. Pelagic production in the Bay over depositional LaHave
clay sediments provided an input of 60gC-m~2-yr~1 from a total
phytoplankton primary production of 133 g C m~2 yr"1 to the benthos.
All of this input could not be utilized by macrofauna or accounted for by
the measured sediment burial rates. It was suggested (Emerson et al.
1986) that the pelagic input, consisting of equal amounts of detritus,
copepod faeces, and microalgae, was tidally transported by tidal excur-
sions of -20 km to a harder sediment on which a productive horse mussel
Modiolus modiolus reef was situated (Table 8.8, number 2). The maxi-
mum production of the macrofauna here was 148gC-m~2-yr"1 at depths
of 60-100 m. Assuming a euphotic depth of 20-40 m, this exceeds the
mean primary production of phytoplankton in the column of seawater
above this part of the lower Bay. It also provides circumstantial evidence
for an alternative source of food by turbulent tidal transport, other than
the direct vertical link between pelagos and benthos proposed by Rowe
(1971). A further example of a partial carbon budget estimate on the
tidal flats of Konigshafen in the German Wadden Sea was provided by
Asmus and Asmus (1990). It also shows the inadequacy of local
autochthonous primary production in the seawater above a blue mussel
reef to provide sufficient carbon sources for the mussel production real-
ized. Estimates of primary production by local phytoplankton amount to
73 g C m~2 yr"1 compared to secondary production of the mussel reef of
264gC-m~2-yr~1. This led Asmus and Asmus (1990) to conclude that the
high suspension feeder production is maintained by tidal transport of
seston from the North Sea or other parts of the Wadden Sea.
348 Benthic Suspension Feeders and Flow
POM POM
COPEPODS CILIATES FLAGELLATES
16
HERRING
Figure 8.12 Partial carbon budget for the Bay of Fundy ecosystem on LaHave
clay sediments at 50-80m (Emerson et al. 1986): annual production, gC-nT2-
yr"1; bracketed figure estimated by the method of Azam et al. (1983).
for this purpose. From tidal velocity measurements made within the
tunnel, the volumetric flow could be calculated and hence the flux of
materials determined. Dame et al. (1984) showed that the oyster reef
reduced particulate organic carbon and chlorophyll a in thefloodingtidal
prism and added significant amounts of excreted ammonia as it ebbed.
Using the benthic ecosystem tunnel, Dame et al. (1992) have determined
the annual flux of ammonia to and from the oyster reef with a net annual
release of 124.8 g-m^-yr"1. In another study, Dame and Libes (1993)
investigated plant nutrient fluxes after experimental manipulations in-
volving removing oysters in some plots. Because the recycling rates of
the oysters were faster than the tidal flushing rates, nutrients were recy-
cled in situ. The controlling feedback loop for nitrogen was suggested
to be the exchange between phytoplankton and oysters, the latter retain-
ing nitrogen within the body.
Field studies on Dutch and German Wadden Sea blue mussel reefs
(Dame and Dankers 1988; Prins and Smaal 1990; Asmus, Asmus, and
Reese 1990; Smaal and Prins 1993) have confirmed hypothesis 4 (Table
8.8): that large quantities of seston are filtered from seawater; faecal and
pseudofaecal biodeposits are produced in large quantities, the latter are
rapidly remineralized, resulting in release of dissolved and inorganic
nutrients in seawater, and this results in stimulation of further
phytoplankton primary production as well as export of nutrients at cer-
tain times of the year. The results are also consistent with H2 of number
3 (Table 8.8), that is that the reef represents a major coupling between
pelagos and benthos.
Nitrogen cycling was investigated by Kaspar et al. (1985) at a green
mussel Perna canaliculus aquaculture site in the Marlborough Sounds,
New Zealand. The main pathways of the nitrogen cycle considered are
shown in Fig. 8.13. Losses of nitrogen occurred when mussels were
removed at harvest and by denitrification processes in sediments and
seawater. Additions occurred from mussel excretion, mainly of ammo-
nia, and the microbial transfer of various organic nitrogen sources which
were present within the sediments. Nitrogen fixation may also have been
important but was not considered in this study. Although it was clear that
there was considerable tidal energy available at the farm, e.g.
Umax = 111 cm-s"1 and a tidal prism of 3-4m, no estimate of tidal trans-
port of nitrogen into or out of the farm area was included in this study.
By comparison with surface sediments at a nearby reference site, signifi-
cantly higher denitrification and nitrogen mineralization rates were
present at the farm.
Ecosystems and Flow 351
Figure 8.13 Nitrogen cycle of a green mussel Perna canaliculus farm, Marl-
borough Sounds, New Zealand (Kaspar et al. 1985).
Summary
Two common features can be discerned in our review of the three
distinct ecosystems we have chosen for inclusion here: kelp forests
seagrass meadows, and bivalve reefs. The first is that there is evidence in
the perturbation theory for the suggestion by Odum et al. (1979) that
ecosystem productivity is a unimodal function of water motion. The
second common feature is that each species of kelp, seagrass, and bivalve
appears to have a characteristic unimodal response to water movement
with distinct tolerance ranges of water movement, as found by Conover
(1968) for some species of kelps and seagrasses. In future work, it should
be possible to utilize this relationship to predict water movement ener-
gies and perhaps also secondary heterotrophic productivity of suspen-
sion feeders.
Ecosystem analysis requires consideration and resolution of the tem-
poral and spatial scales to be used in the study. It is simply not sufficient
for a biologist to set seasonal and geographic limits arbitrarily. This
should be done after a scoping study involving a biologist and a physical
oceanographer. This point of view is forced on us by the examples
reviewed here, which include the South African kelp ecosystem studies
reported by Wulff and Field (1983), the Bay of Fundy horse mussel reef
studies reported in Emerson et al. (1986), and the Wadden Sea blue
mussel reef studies of Asmus and Asmus (1990). These reefs were shown
by Asmus (1994) to depend trophically on turbulent advective transfer of
seston from sources external to the reef subsystem. Recent studies, e.g.
by Grant et al. (1993), use an interdisciplinary approach and ecosystem
354 Benthic Suspension Feeders and Flow
References
Ackerman, J. D. 1983. Current flow around Zostera marina plants and flowers:
implications for submarine pollution. Biol. Bull. 165: 504.
Ackerman, J. D., and A. Okubo. 1993. Reduced mixing in a marine
macrophyte canopy. Funct. Ecol. 7: 305-309.
Ambrose, W. G., Jr., and E. A. Irlandi. 1992. Height of attachment on
seagrass leads to trade-off between growth and survival in the bay scallop
Argopecten irradians. Mar. Ecol. Prog. Ser. 90: 45-51.
Anderson, S. M., and A. C. Charters. 1982. A fluid dynamic study of seawater
flow through Gelidium nudifrons. Limnol. Oceanogr. 27: 399-412.
Asmus, H. 1994. Benthic grazers and suspension feeders: which one assumes
the energetic dominance in Kongigshafen? Helg. Meeresunters. 48: 217-
231.
Asmus, H., and R. M. Asmus. 1990. Trophic relationships in tidal flat areas: to
what extent are tidal flat areas dependent on imported food? Neth. J. Sea
Res. 27: 93-99.
Asmus, H., R. Asmus, and K. Reese. 1990. Exchange processes in an intertidal
mussel bed: a Sylt-flume study in the Wadden Sea. Ber. Biol. Anst.
Helgoland. 6: 1-79.
Azam, F., T. Fenchel, J. G. Field, J. S. Gray, L. A. Meyer-Reil, and F.
Thingstad. 1983. The ecological role of water-column microbes in the sea.
Mar. Ecol. Prog. Ser. 10: 257-263.
Backus, G. J. 1994. Coral reef ecosystems. A.A. Balkema, Rotterdam.
Bernard, F. R. 1974. Annual biodeposition and gross energy budget of mature
Pacific oysters, Crassostrea gigas. J. Fish. Res. Board Can. 31: 185-190.
Bertness, M. D. 1984. Ribbed mussels and Spartina alterniflora production in a
New England salt marsh. Ecology 65: 1794-1807.
Blackburn, T. H., D. B. Nedwell, and W. J. Wiebe. 1994. Active mineral
cycling in a Jamaican seagrass sediment. Mar. Ecol. Prog. Ser. 110: 233-
239.
Butman, C. A., M. Frechette, W. Rockwell Geyer, and V. R. Starczak. 1994.
Flume experiments on food supply to the blue mussel, Mytilus edulis L.,
as a function of boundary layer flow. Limnol. Oceanogr. 39: 1755-1768.
Carlton, J. T., J. K. Thompson, L. E. Schemel, and F. H. Nichols. 1990.
Remarkable invasion of San Francisco Bay (California, USA) by the
Asian clam Potamocorbula amurensis. I. Introduction and dispersal.
Mar. Ecol. Prog. Ser. 66: 81-94.
Cloern, J. E. 1982. Does the benthos control phytoplankton biomass in South
San Francisco Bay? Mar. Ecol. Prog. Ser. 9: 191-202.
Coen, L. D., and K. L. Heck. 1991. The interacting effects of siphon nipping
and habitat on the bivalve (Mercenaria mercenaria (L.)) growth in a
subtropical seagrass {Halodule wrightii Aschers) meadow. J. Exp. Mar.
Biol. Ecol. 145: 1-13.
356 Benthic Suspension Feeders and Flow
Juday, C. 1940. The annual energy budget of an inland lake. Ecology 21: 438-
450.
Judge, M. L., L. D. Coen, and K. L. Heck, Jr. 1993. Does Mercenaria
mercenaria encounter elevated food levels in seagrass beds? Results from
a novel technique to collect suspended food resources. Mar. Ecol. Prog.
Ser. 92: 141-150.
Kaspar, H. F., P. A. Gillespie, I. C. Boyer, and A. L. MacKenzie. 1985. Effects
of mussel aquaculture on the nitrogen cycle and benthic communities in
Kenepuru Sound, Marlborough Sounds, New Zealand. Mar. Biol. 85: 127-
136.
Kenworthy, W. J., J. C. Zeiman, and G. W. Thayer. 1982. Evidence for the
influence of seagrasses on the benthic nitrogen cycle in a coastal plain
estuary near Beaufort, North Carolina (USA). Oecologia 54: 152-158.
Kerswill, C. J. 1949. Effects of water circulation on the growth of quahaugs
and oysters. J. Fish. Res. Board Can. 7: 545-551.
Koehl, M. A. R. 1986. Seaweeds in moving water: form and mechanical
function, p. 603-634. In T. J. Givnish (ed.) On the economy of plant form
and function. Cambridge University Press, Cambridge.
Koehl, M. A. R., and R. S. Alberte. 1988. Flow, flapping and photosynthesis of
Nereocystis lewkeona: a functional comparison of undulate and flat blade
morphologies. Mar. Biol. 99: 435^44.
Koseff, J. R., J. K. Holen, S. G. Monismith, and J. E. Cloern. 1993. Coupled
effects of vertical mixing and benthic grazing on phytoplankton
populations in shallow turbid estuaries. J. Mar. Res. 51: 843-868.
Langton, R. W. 1977. Digestive rhythms in the mussel, Mytilus edulis.
Mar. Biol. 41: 53-58.
Langton, R. W., and P. A. Gabbott. 1974. The tidal rhythm of extracellular
digestion and the response to feeding in Ostrea edulis L. Mar. Biol. 24:
181-187.
Leigh, E. G., R. T. Paine, J. F. Quinn, and T. H. Suchanek. 1987. Wave energy
and intertidal productivity. Proc. Natl. Acad. Aci. U. S. A. 84: 1314-1318.
Lin, J. 1989. Importance in location in the salt marsh and clump size on
growth of ribbed mussels. J. Exp. Mar. Biol. Ecol. 128: 75-86.
Lindeman, R. L. 1942. The trophic-dynamic aspect of ecology. Ecology 23:
399^18.
Loo, L., and R. Rosenberg. 1989. Bivalve suspension-feeding dynamics and
benthic-pelagic coupling in an eutrophicated marine bay. J. Exp. Mar.
Biol. Ecol. 130: 253-276.
Mann, K. H. 1972. Ecological energetics of the seaweed zone in a marine
bay on the Atlantic coast of Canada. II. Productivity of the seaweed.
Mar. Biol. 14: 199-209.
1973. Seaweeds: their productivity and strategy for growth. Science 182:
975.
Mathers, N. F. 1976. The effects of tidal currents on the rhythms of feeding
and digestion in Pecten maximus L. J. Exp. Mar. Biol. Ecol. 24: 271-284.
Miller, R. J., and K. H. Mann. 1973. Ecological energetics of the seaweed zone
in a marine bay on the Atlantic coast of Canada. III. Energy
transformations by sea urchins. Mar. Biol. 18: 99-114.
Menge, B. A., and J. P. Sutherland. 1976. Species diversity gradients: synthesis
of the roles of predation, competition and temporal heterogeneity. Am.
Nat. 110: 351-369.
Monismith, S. G., J. R. Koseff, J. K. Thompson, C. A. O'Riordan, and H. M.
Ecosystems and Flow 359
small fishes and its impact on growth of the bivalve Protothaca staminea
(Conrad). J. Exp. Mar. Biol. Ecol. 63: 249-268.
Peterson, C. H., H. C. Summerson, and P. B. Duncan. 1984. The influence of
seagrass cover on population structure and individual growth rate of a
suspension-feeding bivalve, Mercenaria mercenaria. J. Mar. Res. 42: 123-
138.
Pohle, D. G., V. M. Bricelj, and Z. Garcia-Esquivel. 1991. The eelgrass
canopy: an above-bottom refuge from benthic predators for juvenile bay
scallops Argopecten irradians. Mar. Ecol. Prog. Ser. 74: 47-59.
Prins, T. C, N. Dankers, and A. C. Smaal. 1994. Seasonal variation in the
filtration rates of a semi-natural mussel bed in relation to seston
composition. J. Exp. Mar. Biol. Ecol. 176: 69-88.
Prins, T. C, and A. C. Smaal. 1990. Benthic-pelagic coupling: the release of
inorganic nutrients by an intertidal bed of Mytilus edulis. p. 89-103. In M.
Barnes and R. N. Gibson (eds). 24th European Mar. Biol. Symp.
Aberdeen University Press.
Raupach, M. R., and A. S. Thorn. 1981. Turbulence in and above plant
canopies. Ann. Rev. Fluid. Mech. 13: 97-130.
Reusch, T. B. H., A. R. O. Chapman, and J. P. Groger. 1994. Blue mussels
Mytilus edulis do not interfere with eelgrass Zostera marina but fertilize
shoot growth through biodeposition. Mar. Ecol. Prog. Ser. 108: 205-282.
Rowe, G. T. 1971. Benthic biomass and surface productivity, p. 441-454. In
J. D. Costlow (ed.) Fertility in the sea. Gordon & Breach, New York.
Rowe, G. T., C. H. Clifford, K. L. Smith, and P. C. Hamilton. 1975. Benthic
nutrient regeneration and its coupling to primary productivity in coastal
waters. Nature 255: 215-217.
Schlichting, H. 1968. Boundary-layer theory. 6th ed. J. Kestin (trans.)
McGraw-Hill, New York.
Scoffin, T. P. 1970. The trapping and binding of sub tidal carbonate sediments
by marine vegetation in Bimini Lagoon, Bahamas. J. Sed. Petrol. 40: 249-
273.
Shieh, W. Y., U. Simidu, and Y. Maruyama. 1989. Enumeration of nitrogen-
fixing bacteria in an eelgrass {Zostera marina L.) bed. Microb. Ecol. 18:
249-259.
Smaal, A. C, and T. C. Prins. 1993. The uptake of organic matter and the
release of inorganic nutrients by bivalve suspension feeder beds, p. 271-
298. In R. F. Dame (ed.) Bivalve feeders in estuarine and coastal
ecosystem processes. NATO ASI Ser. Vol. G33. Springer-Verlag, Berlin.
Sorokin, Y. I. 1993. Coral reef ecology. Springer-Verlag, Berlin.
Steever, E. Z., R. S. Warren, and W. A. Niering. 1976. Tidal energy subsidy
and standing crop production of Spartina alterniflora. Est. Coast. Mar. Sci.
4: 473-478.
Thayer, G. W., S. M. Adams, and M. W. LaCroix. 1975. Structural and
functional aspects of a recently established Zostera marina community. In
L. E. Cronin (ed.) Estuarine research. Vol. 1. p. 518-540. Academic Press,
New York.
Tracey, G. A. 1988. Feeding reduction, reproductive failure, and mortality in
Mytilus edulis during the 1985 "brown tide" in Narragansett Bay, Rhode
Island. Mar. Ecol. Prog. Ser. 50: 73-81.
Valiela, L, J. M. Teal, S. Volkman, D. Shafer, and E. J. Carpenter. 1978.
Nutrient and particulate fluxes in a salt marsh ecosystem: tidal exchanges
Ecosystems and Flow 361
and inputs by precipitation and ground water. Limnol. Oceanogr. 23: 798-
812.
Wheeler, W. N. 1980. Effect of boundary layer transport on the fixation of
carbon by the giant kelp Macrocystis pyrifera. Mar. Biol. 56: 103-110.
Wickens, P. A., and J. G. Field. 1986. The effect of water transport on
nitrogen flow through a kelp bed community. S. Afr. J. Mar. Sci. 4: 79-92.
Wolaver, T. G., J. C. Zieman, R. Wetzel, and K. L. Webb. 1983. Tidal
exchange of nitrogen and phosphorus between a mesohaline vegetated
marsh and the surrounding estuary in the lower Chesapeake Bay. Est.
Coast. Shelf Sci. 16: 321-332.
Wulff, F. V., and J. G. Field. 1983. Importance of different trophic pathways in
a nearshore benthic community under upwelling and downwelling
conditions. Mar. Ecol. Prog. Ser. 12: 217-228.
9
Future directions
362
Future Directions 363
Number of
hypotheses Subject Chapter Associated theory
21 Ecosystems Suspension-feeding benthic
communities are important in
pelagic-benthic materials
cycling
15 Larval biology Passive-active theory of
factors influencing the early
life history recruitment of
benthic suspension feeders
Benthic populations Theory of benthic limitation
byflow:trophic group
mutual exclusion; benthic
impoverishment by water
movement
5 Initial suspension- 4 Unimodal theory of
feeding responses suspension feeding to flow,
and bivalve pump theory
4 Behavioral 6 Aggregation theory
responses to flow
0 Filtration physiology 5 Particle collection theory
and ultimate, and, at least until now, our questions were limited to the
former. In this chapter, perhaps because all we are doing is asking and
need not answer them here, we have the luxury of posing both proximate
and ultimate questions.
We believe that there are three important additional steps which have
to be answered affirmatively in selecting a pertinent scientific question:
Is the question of personal interest to the scientist?
Is the question posed testable by scientific methods?
Will the information gained in the answer be relevant?
The first two points seem self-evident. Because of the considerable
amount of work required to complete a research project, an individual
with sufficient motivation and commitment is required to complete it.
We include here the private, moral responsibility of an individual scien-
tist to ascertain, before it is begun, that the research contemplated is not
contrary to the common good. There will always be some questions that
Table 9.2. Judgements on whether alternate hypotheses are upheld by the experimental evidence presented
in Chapters 3-8."
Further
Table No. Hypothesis Yes No tests
3.3 1 Estuarine retention of larvae X X
2 Passive shoreward transport by wind X X
3 Passive shoreward transport by internal waves X X
4 Passive shoreward transport by internal tidal bores X X
5 Passive shoreward transport by upwelling X X
6 Plant canopies entrain larvae X X
a
If further tests are needed as confirmation, this is indicated in the final column.
366 Benthic Suspension Feeders and Flow
Scale
Spatial Temporal
No. (km) (y) Question
0.1-0.01 0.01 Do hydrodynamic-sediment interactions determine the settlement success of deposit versus
suspension feeding larvae so that one or the other is variably excluded from the population by the
physical conditions?
b 0.1-0.01 0.1 Do hydrodynamic-seston interactions determine the recruitment success of suspension feeders at a
particular location?
c 0.1-0.01 0.1 Can we present a model of the internal regulators and external environmental variables which
influence filtering, feeding, and growth of a named suspension feeder?
d What is the mechanism of filtration in active suspension feeders?
e 0.001 0.0 How do tube-living suspension-feeding epifauna respond behaviorally to flow?
f What ethological changes among epifaunal suspension feeders minimize dislodgement by flow?
g 0.001-0.1 >1 What causes epifaunal suspension feeders to aggregate at some sites?
h 0.1 Do hydrodynamic-sediment interactions determine the post-settlement mortality rate of deposit-
versus suspension-feeding populations in such a way that one or the other is variably excluded by
the physical conditions?
i 0.1 >1 What are the mechanisms of water movement-induced inhibition of suspension feeder
production?
i 1-1000 How do physical oceanographic currents and wind-wave effects interact to influence pelagic-
benthic coupling in a macrotidal estuary?
Future Directions 369
For the two ethological questions of Table 9.3 (e and f), we believe
that both field and laboratory observations of a wide range of species of
suspension feeder are required. For locations which are deeper than
SCUBA diving depth, underwater, time-lapse photography, from either
an anchored vessel or a remotely operated vehicle, may be adequate. In
the laboratory, a video camera focused through the acrylic plastic
(Plexiglas) walls of the flume may often be suitable. With the latter
equipment, it should be possible to record tube-living suspension feeder
behavior as a function of unidirectional or oscillatory flows - some
results with spionid polychaetes were discussed in Chapter 6. As postu-
lated in Chapter 4, we would expect feeding to be unimodally related to
velocity, and at the upper end of this relationship, to see behavioral
responses to minimize dislodgement by water movement, e.g. halting
feeding and retiring within the tube. Free-living epifauna may also re-
spond in a characteristic way toflowswhich are nearly strong enough to
sweep them away. We would expect that each species could tolerate a
particular velocity range and thus would indicate to the experienced
benthic biologist which ambient flows were available at the site. The
bottom line is that more detailed observations of suspension feeder
behavior must be completed before hypotheses and experimentation can
be initiated to determine the mechanisms involved.
In some ways, question (i) of Table 9.3 is a continuation of (f), al-
though it is cast at the synecological level. Numerous studies have shown
the drastic influence that a violent storm can have on nearshore benthic
communities (see Chap. 7). Even if it were possible to simulate wave
erosion events in the flume, it would be necessary to provide natural
groups of organisms, since it has been established (Chap. 6) that the
erosion threshold may be influenced by the presence of other individuals
of the same, or different, species. Almost certainly there will be energetic
costs associated with maintaining position in flows, whether behavioral,
e.g. in the ability of sand dollars to burrow prior to storms and scallops to
recess into sediments (Chap. 6), or structural, e.g. in skeletal stiffening by
spicule formation in sponges, deposition of calcium salts in corals, or
secretion of adhesives by barnacles (Chap. 6). As far as we are aware, the
comparative energy costs associated with this effort have yet to be deter-
mined. Such costs would be expected to increase with the water move-
ment experienced by populations at each locality examined.
At subcritical erosion thresholds in either unidirectional or oscillating
flows, there is an inhibitory effect on feeding - the c stage of Chapter 4
Future Directions 373
Scale
Spatial Temporal
No. (km) (yr) Question
0.001 ~1 What are the environmental variables which
affect the physiology of feeding/growth in
commercially important bivalves?
1.0 ~1 What is the carrying capacity of blue mussels in
suspension culture in a named location?
10-100 1-100 What are the benthic and consequent ecosystem
effects of landfilling and wharf building at the
mouth of an estuary?
1-1000 1-100 What are the near and far field ecological effects
of building a major tidal power dam?
Ultimate questions
The history of life on planet Earth is considered by Lincoln, Boxshall,
and Clark (1982) to have begun some 4500 x 106yr ago. Throughout this
time, tectonic plate movements of all the major land masses and eustatic
changes in seawater levels have repeatedly occurred (Myers 1994). Cli-
matic and ocean current changes separate from or associated with these
events have also occurred (Berggren and Hollister 1977). Thus, the
evidence from geology suggests continuous environmental change
through geological time caused by these universal forcing functions.
Additional discontinuous events such as meteorite impacts (Swinburne
1993) or violent earthquake activity (Officer 1993) may have sudden and
catastrophic effects on animal life on both sea and land. The importance
of violent discontinuous events in relation to extinction events is still
being debated (e.g. Moses 1989; Quinn and Signor 1989; Serjeant 1990).
Because of the adaptational nature of plant and animal life, we should
expect continuous changes - evolution or extinction - of these organisms
to fit, or exclude, them from such changing environments.
All of the theories or hypotheses which can be formulated for the
ultimate questions we have posed in Table 9.6 can only be addressed by
deductive methods. In deduction, the formal scientific propositions must
be made without direct observation of the natural or experimentally
contrived events involved. Because the scientist cannot be present when
the evolutionary events associated with the questions in Table 9.6 occur,
the only evidence for, or against, an hypothesis is circumstantial. A
frequent problem with circumstantial evidence is knowing how much is
required to prove that a particular hypothesis is correct or to choose
between competing hypotheses when the balance of evidence is nearly
Future Directions 379
a
From Table 1.2.
equal between them. Thus, deductive methods lack the power of deci-
sion available to those answering proximate questions using inductive
methods.
As background to our consideration of ultimate questions (Table 9.6),
we discuss three common types of circumstantial evidence which might
be available. The first involves the geological record and could include
direct preservation of parts of the organism, which gives clues to the
species evolutionary history. Alternatively, the geological record could
provide inferences about past climates and environment from the types
of organisms preserved in ancient sediments, e.g. the palaeoecology of
marine sediments (Molfino 1994). Surficial geologists study modern
sediments and the benthic flora and fauna they contain to improve their
view of the factors which may have operated in the geological past (e.g.
Rhoads and Young 1970; Thorns and Berg 1985). Surficial geologists also
have independent means of measuring the age of sediments by isotopic
decay rates (Molfino 1994). A second general approach is through
phylogeny. Modern taxonomy requires determining a wide range of,
largely, morphological characters in each species, and this has led to the
380 Benthic Suspension Feeders and Flow
References
Amos, C. L., and D. A. Greenberg. 1980. The simulation of suspended
participate matter in the Minas Basin, Bay of Fundy: a region of potential
tidal power development, p. 2-20. In Proceedings Canadian Coastal
Conference, Nat. Res. Counc, Ottawa.
Bacher, C, M. Heral, J. M. Deslous-Paoli, and D. Razet. 1991. Modele
energetique unibiote de la croissance des huitres (Crassostrea gigas) dans
le bassin de Marennes-Oleron. Can. J. Fish. Aquat. Sci. 48: 391-404.
Baker, G. C. 1984. Engineering description and physical impacts of the most
probable tidal power prospect(s) under consideration for the upper
reaches of the Bay of Fundy, p. 333-345. In D. C. Gordon and M. J.
Dads well (eds.) Update on the marine environmental consequences of
tidal power development in the upper reaches of the Bay of Fundy.
Can. Tech. Rep. Fish. Aquat. Sci. 1256.
Bayne, B. L., R. J. Thompson, and J. Widdows. 1976. Physiology I, p. 121-206.
388 Benthic Suspension Feeders and Flow
391
392 Glossary
lected to the rate at which they would pass through the space occu-
pied by the collector if it were not there.
entrainment the physical process of transfer of fluid by friction from
one water mass to another. Usually occurs between currents moving
in different directions.
eustatic the slow, universal changes of sea level which are caused by the
melting of continental ice sheets.
fall velocity the rate at which a body passively falls under the influence
of gravity through a still fluid; same as settling velocity.
Fick's laws of molecular diffusion the first law states that the flux of
molecules is equal to the molecular diffusion coefficient times the
concentration gradient. The second law relates the rate of change of
concentration at a specified time and position to the diffusion coef-
ficient and the second derivative of concentration with respect to
distance.
field flume a flume capable of being used in the field, either in situ, as
by surficial geologists, or by benthic biologists where the flume is
portable but not necessarily used in situ.
first law of thermodynamics the conservation of energy law: energy
may be transformed from one form to another, but cannot be cre-
ated or destroyed.
flow cytometer a device which uses optical analysis based on light
scattering and fluorescence for measuring the size and epi-
fluorescence properties of small particles, e.g. red blood cells or
seston.
flow fluorometer a device which measures the fluorescent radiation of
a sample after exposure to monochromatic radiation. The sample is
continuously brought to the sensor in a pumped flow of seawater,
e.g. continuous seawater sampling of chlorophyll a in living phyto-
plankton cells.
flow straightener a collimator (q.v.).
flow meter a device for measuring the volumetric flow rate of a fluid.
flow vector the direction and speed of aflowat a given point in a fluid.
flowthrough flume a type of flow channel where the water makes only
a single pass through the experimental working section.
fluid a liquid or a gas.
flume boundary layer the sheared layer that develops on the side walls
and bottom of flumes during passage of water.
force the external agency which changes the state of rest or motion of
a body and which is measured in units of dynes in the cm/g/s system.
396 Glossary
critical force. For this force the static friction is the difference be-
tween buoyant and gravitational forces on the body, multiplied by
the static friction coefficient.
steady flow a flow field which does not change with time.
stochastic a system model in which the relationships are random or
probabilistic, so that an input results in many possible outcomes
decided by chance. Opposite of deterministic, in which an input
results in one outcome only.
Stokes law the relation of the total drag, F, on a sphere of diameter, D,
moving in a still Newtonian fluid, to its velocity, [/, and the viscosity
of the fluid, (i, in creeping flow conditions. Thus: F=3n[iDU.
strain gauge a device for measuring the force of attachment of
an organism to a hard substrate, by the use of attached pressure
sensors which change in electrical resistance as the sensor is
strained.
stratification the separation of a water body into discrete horizontal
layers of the water column by physical environmental variables, e.g.
salinity or temperature.
streamlines the characteristic flow lines of a steady flow as can be
visualized by observing paths of neutrally buoyant particles. The
streamlines are everywhere tangent to the local velocity vector.
Strouhal number a dimensionless frequency used in analyzing oscillat-
ing fluid flows. The ratio of the frequency of the oscillation to
UD'1, where U is the velocity and D a dimension used to character-
ize the flow.
stuw zone the most inland part of some estuaries where freshwater is
backed up by the tidal stream during a flooding tide, causing a
freshwater tidal rise and fall.
submarine canyon a sharply shelving valley on the seabed.
surf zone that part of the shore between the landward limit of wave
uprush and the most seaward breaker.
surface slick that part of the sea surface where up- or down-welling
occurs and which is often marked by flotsam and jetsam, e.g. where
Langmuir cells meet.
surficial geology the scientific study of unconsolidated sedimentary
deposits on the land or seabed.
suspension feeder-biased sampling simulation of feeding by a named
suspension feeder. Sampling may be isokinetic or non-isokinetic,
but at the same velocity as used naturally by the suspension
feeder.
402 Glossary
Specific plant or animal names and words defined in the glossary are excluded.
405
406 Index