You are on page 1of 30

Journal of Fish Biology (2004) 65, 1193–1222

doi:10.1111/j.1095-8649.2004.00568.x, available online at http://www.blackwell-synergy.com

REVIEW PAPER
Fish functional design and swimming performance
R. W. B L A K E
Department of Zoology, University of British Columbia, Vancouver, British Columbia,
V6T 1Z4, Canada

(Received 25 February 2004, Accepted 24 August 2004)

Classifications of fish swimming are reviewed as a prelude to discussing functional design and
performance in an ecological context. Webb (1984a, 1998) classified fishes based on body shape
and locomotor mode into three basic categories: body and caudal fin (BCF) periodic, BCF
transient (fast-starts, turns) and median and paired fin (MPF) swimmers. Swimming perform-
ance and functional design is discussed for each of these categories. Webb hypothesized that
specialization in any given category would limit performance in any other. For example, routine
MPF swimmers should be penalized in BCF transient (fast-start propulsion). Recent studies
offer much support for Webb’s construct but also suggest some necessary amendments. In
particular, design and performance compromises for different swimming modes are associated
with fish that employ the same propulsor for more than one task (coupled, e.g. the same
propulsor for routine steady swimming and fast-starts). For example, pike (BCF transient
specialist) achieve better acceleration performance than trout (generalist). Pike steady (BCF
periodic) performance, however, is inferior to that of trout. Fish that employ different propul-
sors for different tasks (decoupled, e.g. MPF propulsion for low-speed routine swimming and
BCF motions for fast-starts) do not show serious performance compromises. For example,
certain MPF low-speed swimmers show comparable fast-start performance to BCF forms.
Arguably, the evolution of decoupled locomotor systems was a major factor underlying the
adaptive radiation of teleosts. Low-speed routine propulsion releases MPF swimmers from the
morphological constraints imposed by streamlining allowing for a high degree of variability in
form. This contrasts with BCF periodic swimming specialists where representatives of four
vertebrate classes show evolutionary convergence on a single, optimal ‘thunniform’ design.
However, recent experimental studies on the comparative performance of carangiform and
thunniform swimmers contradict some of the predictions of hydromechanical models. This is
addressed in regard to the swimming performance, energetics and muscle physiology of tuna.
The concept of gait is reviewed in the context of coupled and decoupled locomotor systems.
Biomimetic approaches to the development of Autonomous Underwater Vehicles have given a
new context and impetus to research and this is discussed in relation to current views of fish
functional design and swimming performance. Suggestions are made for possible future
research directions. # 2004 The Fisheries Society of the British Isles

Key words: Autonomous Underwater Vehicles; coupled and decoupled locomotor mechanisms;
fish swimming; functional design; gaits; performance.

Tel.: þ1 604 8223373; fax: þ1 604 8222416; email: blake@zoology.ubc.ca

1193
# 2004 The Fisheries Society of the British Isles
1194 R. W. BLAKE

INTRODUCTION
The development of hydromechanical models of aquatic locomotion during the
last 35 years has provided the basis and impetus for some important develop-
ments in organismal biology. For example, Lighthill (1969) conveyed the utility
of hydrodynamic analysis in understanding the morphological designs and
movement patterns of aquatic animals. His seminal review suggested a number
of areas for hydromechanical and biological research that included hydro-
mechanical studies of the fastest aquatic animals (tunas and dolphins), propul-
sion by undulations in a fin alone, propulsion in skates and rays (Rajidae) and
the turning and starting mechanisms of fish. Sparenberg (2002) provides a
through review of these and other mathematical aspects of fish locomotion
research, while the present paper focuses on the principal results of hydro-
mechanical and experimental work on fish functional design and swimming
performance with an emphasis on behavioural and ecological perspectives.
The present review considers the significance of the results (as opposed to
methodology and analysis) of mechanical and biological research for: (1) fish
fast-start swimming, (2) steady propulsion, (3) appendage-based slow swimming
in complex environments. The implications of morphological and locomotor
specializations are discussed in the context of Webb’s (1984a, 1998) construct of
performance trade-offs and the concept of swimming gaits. Biomimetic
approaches towards the development of autonomous underwater vehicles are
also considered.

SWIMMING MODES, CATEGORIES AND STYLES


Classifications of swimming modes based on considerations of the structures
employed, their kinematics and mechanics (axial v. appendage propulsors),
categories based on activity level and duration (continuous cruising v. transient
swimming) and styles (overall movement patterns, e.g. intermittent v. continu-
ous propulsion) are central to the description of the diversity of fish locomotion.
Breder (1926) classified the propulsive movements (modes) of fish based on
‘type’ genera employing the suffix ‘iform’ (e.g. anguilliform: based on the eel
Anguilla). The classification focuses on average expressions of a continuum and
implies nothing about taxonomic or evolutionary relationships. Nevertheless its
use persists, albeit for reasons of convenience only. Lindsey (1978) and Webb &
Blake (1985) have expanded and modified Breder’s classification. Axial undu-
latory propulsors in steady (periodic) continuous swimming fall into four
groups: anguilliform, sub-carangiform, carangiform and thunniform, repre-
senting a series from serpentine undulation (eels, hagfish, certain sharks) to
localized posterior oscillation (tuna). Among those fish not propelled by bodily
and caudal fin undulations, six undulatory fin modes may be distinguished
(amiiform, gymnotiform, balistiform, diodontiform, tetradontiform and
rajiform). Appendage oscillatory propulsors are divided into two types: rowing
(drag-based) and lifting (wing-based) fins.
The different modes of body and caudal fin steady (periodic BCF) and
unsteady (fast-start, BCF transient) propulsion involve different mechanical
principles (Lighthill, 1975; Blake, 1983a; Videler, 1993). Other aspects of the

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1195

classification, however, lack any functional basis. For example, the distinctions
between the various ‘non-body’ undulatory median and/or paired fin modes is
simply based on the number and location of propulsive fins and the general
appearance of the waveforms present. Blake (1980a, 1983a, b) suggested a
simple functional classification based on two groups: Type I; forms that exhibit
waveforms of relatively high amplitude, low frequency and large wavelength
(e.g. Diodontidae): Type II; forms characterized by fins of low amplitude, high
frequency and small wavelength (e.g. Sygnathidae). The kinematic differences
underlying the two groups are reflected functionally in differences in hydro-
mechanical efficiency (Blake, 1983a). The lack of functional relevance regarding
swimming performance of the established classification system is not limited to
fish propelled by undulatory median and/or paired fins. Anguilliform swimmers
show substantial kinematic and morphological variation across taxa which is
reflected in different levels of hydromechanical efficiency (Gillis, 1997). Some
carangiform swimmers show kinematic differences relative to thunniform swim-
mers but no significant differences in maximum sustainable speeds and cost of
transport (Donley & Dickson, 2000; Sepulveda & Dickson, 2000).
Webb (1984a) defines four functional categories of locomotion for aquatic
vertebrates: body and caudal fin (BCF) periodic (steady) propulsion, BCF
transient (fast-starts, powered turns) propulsion, median and paired fin (MPF)
propulsion and occasional or ‘non-swimming’. Two characteristics linked to
feeding (distribution of food in space and time and evasive capabilities in
predator-prey interactions) are correlated with these locomotor categories.
Webb (1984a) points out that BCF periodic swimmers take food that is widely
dispersed, whereas BCF transient forms consume locally abundant evasive prey
and MPF swimmers eat relatively non-evasive food in structurally complex
habitats.
Arguably, the non-functionally oriented, descriptive classification stemming
from Breder (1926) could be abandoned in favour of a functionally based
system consisting of undulatory BCF periodic (steady), BCF transient
(unsteady), undulatory MPF (Type I and II) swimming and oscillatory drag
and lift-based propulsion. Sfakiotakis et al. (1999) employ a similar approach in
a review of fish swimming modes directed at encouraging engineering perspec-
tives on the functional design of Automated Underwater Vehicles (AUVs) based
on fish ‘models’. Nevertheless, given the current continued reference to, and
indeed, continuing extension of ‘Breder style’ classifications [e.g. Fricke and
Hissmann (1992) refer collectively to steady, transient and MPF swimming in
the coelacanth Latimeria as being ‘Coelacanthiform’], reference is made to them
here.
Beamish (1978) classified fish activity levels into three major categories:
sustained, prolonged and burst swimming. Sustained swimming refers to speeds
that can be maintained for long periods (>200 min) without muscular fatigue.
Prolonged swimming (20 s to 200 min) ends in fatigue. Burst swimming ends in
fatigue in <20 s and corresponds to values of c. 10 body lengths per second
(bl s1) for sub-carangiform fish of between c. 10 and 20 cm in length
(Bainbridge, 1960). The concept of critical swimming speed (Brett, 1964) is
often employed to designate the maximum velocity that a fish can maintain
for a precise time period. Critical swimming speeds do not necessarily give

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1196 R. W. BLAKE

accurate estimates of endurance in prolonged swimming. Farlinger & Beamish


(1977) have shown that endurance is influenced by the time increments selected
in critical swimming speed determinations. Kolok (1992) suggests that ‘step-
based’ critical swimming determinations may pre-exhaust fish to anaerobic
levels. Lee et al. (2003a) determined that the anaerobic cost of swimming to
critical velocity in sockeye salmon Oncorhynchus nerka (Walbaum) and coho
salmon Oncorhynchus kisutch (Walbaum) can be up to c. 50% of the measured
oxygen consumption.
Some studies employ less formalized methods focused on assessing a par-
ticular level of swimming performance based on a fixed (or ‘fatigue’) velocity.
For example, Sepulveda & Dickson (2000) determine the maximum sustainable
speed in juvenile Kawakawa tuna Euthynnus affinis Cantor and chub mackerel
Scomber japonicus Houttuyn based on an assessment of the maximum speed
that a fish can maintain swimming against a current in a flume with continuous
tail beats employing slow twitch oxidative muscle for a 30 min period. The
criteria for defining the maximum sustainable speed was when a fish switched
to burst and glide swimming three times within a 30 s period implying the
recruitment of glycolytic muscle fibres (Rome, 1995). In addition, results
obtained from incremental and fixed velocity protocols are not necessarily
directly comparable (Hammer, 1995).
Most determinations of swimming performance are based on forcing fish to
swim in a flume against a unidirectional current. This raises the question of the
extent to which laboratory estimates are relevant to natural circumstances
where fish are free to select their movement style and activity level (Blake,
1991). Boisclair & Tancy (1993) assessed the energy cost (as oxygen consump-
tion rate) of different species of fish forced to swim against a unidirectional
constant current, in stationary water and swimming routinely (changes in speed
and direction). They found that routine swimming was most expensive, followed
by quasi-steady swimming in stationary water and forced swimming against a
current. To be ecologically relevant, assessments of swimming energetics in fish
and other aquatic animals must take account of natural swimming behaviours.
Hughes & Kelly (1996) describe an effective technique based on two or more
video cameras for tracking swimming fish in three dimensions in still or flowing
water, allowing for an indirect biomechanical assessment of swimming costs.
Lee et al. (2003b), however, show that accurate respirometric measurements can
be made in the field employing large portable flumes. Lee et al. (2003b) studied
different British Columbia stocks of sockeye and coho salmon over their natural
ambient temperature ranges (5–20 C) and ‘in-river’ migration distances
(ranging from c. 100 to 1000 km for coastal and interior stocks respectively).
In addition to expected temperature effects on metabolic rate, Lee et al. (2003b)
found significantly higher optimal temperatures for the maximum rate of
oxygen consumption, scope for activity and cost of transport for the interior
stocks (longer migrations) relative to the coastal stocks (shorter migrations).
Intraspecific differences in swimming performance in the context of different
life-history strategies are common in fish. Taylor & Foote (1991) found intra-
specific differences in endurance between sockeye salmon (anadromous) and
non-anadromous kokanee and Taylor & McPhail (1986) have documented
differences in the prolonged swimming performance of anadromous and

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1197

non-anadromous three-spined sticklebacks Gasterosteus aculeatus L. In both


cases the performance of the anadromous fish was superior to that of the
non-anadromous form.
In addition to exploring intraspecific life-history differences, comparative
determinations of swimming performance have been employed to quantify the
functional significance of body form differences in the context of intraspecific
trophic polymorphisms associated with differences in habitat use and the adap-
tive radiation of evolutionary lineages. Salmonids frequently exhibit different
sympatric forms that differ in body size, gill raker number and other morpho-
logical traits linked to benthic v. open-water habitats (Behnke, 1972; Skulason
et al., 1989; Jonsson & Jonsson, 2001). Ehlinger & Wilson (1988) and Ehlinger
(1990) described the adaptive intraspecific variation in morphology and
foraging behaviours in bluegill sunfish (Lepomis macrochirus Rafinesque) and
demonstrated that a littoral form of L. macrochirus is characterized by longer
periods of hovering and larger pectoral fins than an open-water limnetic
form. Law & Blake (1996) document differences in swimming performance
associated with the adaptive radiation of two morphologically distinct limnetic
and benthic lake-dwelling species of three-spined sticklebacks. The adaptive
radiation of Gasterosteus spp. in the coastal lakes of British Columbia is
characterized by selective pressure for differential sustained and prolonged
swimming performance (limnetics superior to benthics) linked to feeding and
a conservation of fast-start escape capability (associated with similar predation
pressure).
Distinct from locomotor mode and activity category are swimming styles.
Examples include burst and glide and burst and coast swimming in negatively
and neutrally buoyant fish respectively, tidal-stream transport and ground effect
in benthic negatively buoyant fish. The energy requirements of these swimming
patterns are significantly less than those for steady rectilinear propulsion (see
Blake, 1983a, 2000; for reviews).

FAST-STARTS, TURNS, FUNCTIONAL DESIGN AND PISCIVOROUS


PREDATOR-PREY INTERACTIONS
Fast-start (BCF periodic) swimming is central to understanding the dynamics
of fish predator-prey interactions because most fish employ them to escape from
predators and to capture prey. Indeed, fast-start swimming is a characteristic
feature of fish evolution, from agnathans to crossopterygians. O’Steen et al.
(2002) have shown that fast-start behaviour is closely linked to survival and
evolves quickly with changes in predation pressure. Fish that do not execute
fast-starts depend on other mechanisms to deter and avoid predators
(e.g. armor, spines, poisons, camouflage), and highly specialized prey-capture
mechanisms (e.g. lures).
Most biomechanical studies of fast-start swimming focus on BCF swimmers
and consider performance levels (e.g. distance covered, velocity, acceleration) in
relation to body form (Hertel, 1966; Weihs, 1973a; Webb, 1977, 1978a, b;
Harper & Blake, 1990; Frith & Blake, 1991, 1995). Early studies (Gero, 1952;
Hertel, 1966; Weihs, 1973a) report maximum accelerations of 40–50 m s2 and
describe fast-start kinematics. Blake (1983a) reviews the basic kinematics of fish

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1198 R. W. BLAKE

fast-start swimming. Essentially, there are two basic types defined relative to the
shape of the body during the event (C- and S-starts), and each type involves
three phases. In stage 1 the axial musculature bends the fish into a C or S shape,
a contralateral contraction initiates the main propulsive stroke (stage 2) and a
final stage (3) may involve continued active swimming or passive coasting.
C-starts are typically used by escaping prey. Large unbalanced side-forces are
produced during stage 1 of the escape response (C-start) causing the centre of
mass to rotate and turn the body as it moves forward. In stage 2 the propulsive
force is directed more directly backward and the fish accelerates forward more
rapidly. Due to the initial rotation of the head the animal also, however,
continues to turn. The turning rate is a function of the acceleration which
tends to be maximal. S-starts are typical of predator attacks from a standing
start (Webb & Skadsen, 1980). However, some fish employ S-starts for both prey
capture and escape responses. Schriefer & Hale (2004) measured muscle activity
and body kinematics in pike Esox lucius L. prey strikes and startle responses
and show that the movement patterns are similar in both behaviours. Stage
1 (pre-propulsive positioning stage), however, is slower and more variable during
prey strikes and involves less rostral movement than for startle responses.
Schriefer & Hale (2004) suggest that strike behaviour is mediated by more
complex neural circuits than the Mauthner system initiated C-start, allowing
for greater modulation of stage 1 while maintaining high stage 2 performance
for prey strikes. The S-shaped body posture allows the lateral forces generated
to be more balanced along the length of the fish, eliminating uncontrolled
turning as the fish accelerates towards its prey. Later studies have focused on
more specific aspects of fast-starts, such as predator and prey latency periods
(Eaton et al., 1977; Webb, 1984b; respectively), size effects (Webb, 1976; Domenici
& Blake, 1993a), the effects of median fin amputation (Webb, 1977), temperature
effects (Webb, 1978a) and the influence of body form on performance (Webb,
1978b; Harper & Blake, 1990). By the mid 1980s, it was thought that mean and
maximum acceleration rates of 40–50 m s2 were typical and that performance was
essentially independent of body form for fusiform BCF swimmers.
Studies employing critical evaluation of the possible errors involved in
analysing high-speed cinematography and videography to determine maximum
acceleration rates and other experimental approaches (involving subcutaneously
implanted miniature accelerometers) have revised estimates of acceleration
performance and views on the relationship between fast-start performance and
body form. Harper & Blake (1989a, b) analysed film-derived and accelerometer
measured acceleration-time data of fish fast-starts and showed that significant
errors are likely in film analysis when framing rates and image magnification
depart from optimum values. Direct measurements of acceleration in
pike indicated that previous assessments had underestimated performance levels.
Values for maximum acceleration rate during a C-type fast-start of c. 250 m s2
were recorded. This value is about twice as high as previously reported values for
any fish. Knowing the limits to fast-start swimming performance is important for
understanding the possible outcomes of given piscivorous predator-prey inter-
actions. Whilst understanding the limits to fish fast-start performance from the
standpoint of functional design, behaviour and physiology requires more than
knowledge of acceleration performance, it is nevertheless a key ingredient.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1199

Webb (1978b) determined the escape fast-start performance of seven teleost


species of different body form. He compared the performance of fish of differ-
ent lateral profiles because, among other things, hydromechanical theory pre-
dicts that good fast-start performance is enhanced by lateral profiles that
maximize the mass of water accelerated by the body movements (Weihs,
1973a). The results showed little variation in both kinematics and acceleration
performance. Webb (1978b) concluded that fast-start performance depends
primarily on a compromise between muscle mass as a percentage of body
mass and body and fin profile. Although a large depth of section maximizes
the added mass (sum of fish mass and the water entrained), it also reduces the
volume available for the axial white muscle that powers the process.
Harper & Blake (1990) combined cinematography and accelerometry to
assess the escape performances of rainbow trout and pike to test the hypothesis
that mean and maximum accelerations during fast-starts are independent of
body form. Comparisons of distance, time, mean and maximum velocity and
mean and maximum acceleration rates showed that fast-start of pike is signifi-
cantly higher than that of trout for all performance parameters measured.
Harper & Blake (1990) showed that escape fast-start performance is related to
body form. One reason for the disparity between the results of Harper & Blake
(1990) and those of previous studies has been noted (framing rate and film
analysis procedures). The results of Harper & Blake (1990) support previous
suggestions (Weihs, 1973a; Lighthill, 1975; Webb, 1986) that the body form
of pike is well-designed for BCF transient swimming and that of rainbow trout
a compromise, showing some features that enhance BCF periodic (steady)
swimming and some that benefit fast-start performance.
Webb (1986) investigated the effect of body form on the vulnerability of four
species of teleost prey (Esox spp., L. macrochirus, Micropterus salmoides
Lacépède and Pimephales promelas Rafinesque) to predation by M. salmoides.
Esox spp. showed the highest mean acceleration rates followed by L. macrochirus,
M. salmoides and P. promelas, showing that predator induced escape fast-start
performance is different for species of different body form. Other differences
between the species, such as response thresholds, however, were also found to
be important. Harper & Blake (1991) compared the performance of pike in
prey capture to that in escapes. Mean and maximum acceleration and velocity
were higher during escapes than during prey capture showing that fast-start
performance level depends on the purpose. In addition, prey capture perform-
ance in pike is also related to prey size, apparent prey size (defined as the angular
size of the prey on the pike’s retina) and strike distance (the distance from the pike
to the prey at the onset of the fast-start) (Harper & Blake, 1991). Spierts & van
Leeuwen (1999) found that carp Cyprinus carpio L. C-start performance exceeded
that of S-starts. Not all aspects of fast-start performance in fish can be explained on
the basis of form. Wakeling & Johnston (1999) investigated the fast-start perform-
ance during ontogeny of carp and found that both propulsive muscle force and
power increase with size. The hydrodynamic efficiency (relating the mechanical
power production to the inertial swimming power), however, showed no significant
increase during development. Wakeling et al. (1999) concluded that the increased
hydrodynamic power requirements with increasing size are met by increased muscle
power. Law & Blake (1996) compare the prolonged and fast-start swimming

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1200 R. W. BLAKE

performance of a morphologically distinct sympatric species pair (limnetic and


benthic) of lake-dwelling sticklebacks Gasterosteus spp. The ‘limnetics’ are stream-
lined and have a significantly lower drag coefficient and superior prolonged swim-
ming performance than the larger, deeper bodied ‘benthics’. Based on
morphological considerations, it was hypothesized that the fast-start performance
of the benthics would exceed that of the limnetics. However, there was no signifi-
cant difference for any linear or angular parameter of fast-start performance
between the two species. Taylor & McPhail (1986) have documented differences
in prolonged and fast-start performance between anadromous and freshwater
three-spined sticklebacks. Domenici & Blake (1997) give a detailed review of the
kinematics and performance of escape and feeding fast-starts in relation to habitat
and form.
The kinematics and mechanics of turns and fast-starts are similar. Essentially,
C-type fast-starts are turns initiated from a standing start as opposed to being
executed with some forward velocity. Many fish that pursue evasive prey
execute turns whilst chasing and capturing their quarry. Consequently, an
understanding of turning dynamics and performance is central to understanding
many piscivorous predator-prey interactions. Among other things the relative
maneuverability of the predator and prey is an important determinant of the
outcome of predator-prey interactions (Howland, 1974). In a piscivorous
predator-prey interaction, as the prey attempts to escape a pursuing predator
by turning on a circular arc, it experiences a sideways acceleration (velocity
squared divided by the turning radius) towards the centre of the circle. Howland
(1974) showed that if the prey’s sideways acceleration is greater than that of the
predator’s the prey can escape (even if the predator’s top speed is higher) if the
prey swerves when the predator is close behind it.
A common approach to assessing the performance of fish in the context of
predator-prey relations is to determine the minimum turning radius of the
predator and prey in question (Webb, 1983). Unlike fast-starts, swimming
turns are not stereotypic behaviours, among other things, individual motivation
and interactions with other fish may influence performance levels. In addition,
the experimenter has little control over events and given performance levels
cannot be repeated exactly. Nevertheless, specific turning radius (turning radius
divided by body length) is a useful indicator of relative maneuverability. Many
authors report a minimum turning radius based on a number of turning events
(Webb & Keyes, 1981; Webb, 1983; Webb & de Buffrenil, 1990). In practice,
turns of any radius down to the minimum are possible and the average of any
series of measurements likely simply represents the lower (tighter) end of a
range.
Blake et al. (1995) investigated the turning radius of yellowfin tuna Thunnus
albacares (Bonnaterre) and found that the average specific turning radius of
T. albacares (047 body length, L) is large relative to other fish [013 L and
023 L for dolphin fish Coryphaena hippurus L. and yellowtail Seriola dorsalis
Cuvier respectively (Webb & Keyes, 1981), 013 L and 018 L for smallmouth
bass Micropterus dolomieu Lacépède and rainbow trout respectively (Webb,
1983) and 0065 L for angelfish Pterophyllum eimekei Ahl (Domenici & Blake,
1991)]. This supports Webb’s (1984a) hypothesis of mutually exclusive
specialized classes in locomotor design and performance. Yellowfin tuna are

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1201

specialized BCF periodic swimmers and perform poorly in turning relative to


locomotor generalists (e.g. rainbow trout) and MPF specialists (e.g. angelfish).
Data for rainbow trout (Webb, 1976) and angelfish (Domenici & Blake, 1991)
suggest that specific turning radius is constant for a given fish species, implying
that the relative maneuverability of a given piscivorous predator and prey is
too. Whilst the relative maneuverability of predator and prey is a key factor,
however, in the dynamics of predator-prey interactions it is not the only one.
Domenici & Blake (1991) have shown that angelfish show two kinematically
different types of escape involving trade-offs between velocity and turning
angle. Kasapi et al. (1993) argue that angular kinematic parameters
(roll, pitch and turning angle) are important in the escape response of knifefish,
Xenomystus nigri Gunther. Turning performance also depends on body
flexibility. Breinerd & Patek (1998) show that the maximum body curvature
with the first half tail beat of four different species of tetradontiform fishes
during escape swimming is a function of the number of inter-vertebral joints in
the spine. Other important factors include response latency (Eaton & Hackett,
1984), reaction distance (Dill, 1974; Harper & Blake, 1988, 1991), initial
orientation of the prey relative to the predator (Domenici & Blake, 1993b)
and sensory responsiveness (Blaxter & Fuiman, 1990). The kinematics and
behaviour of fish fast-starts in the context of predator-prey relationships are
reviewed by Batty & Domenici (2000) and Domenici (2001) discusses scaling
considerations in the same context.

STEADY SWIMMING, FUNCTIONAL DESIGN AND FEEDING


Optimal morphologies for steady (periodic) BCF propulsion have been iden-
tified from hydromechanical theory (Wu, 1971; Lighthill, 1975; Webb, 1975,
1984a; Blake, 1983a) and addressed experimentally (Webb, 1975, 1988; Weihs,
1989; Videler, 1993). Essentially, the optimal functional design for BCF periodic
propulsion reflects those features that maximize thrust and minimize drag and
are typical of the tunas and their relatives. Thrust is maximized by a high aspect
ratio (large span and relatively small chord) lunate caudal fin, a narrow caudal
peduncle (allowing for large amplitude body movements posteriorly and control
over the angle of attack of the caudal fin), and a large anterior body depth and
mass to minimize recoil energy losses. These features, together with a stream-
lined, bullet-shaped body of near optimal fineness ratio (ratio of body length to
diameter) also act to minimize drag. It has long been known that the drag of
swimming streamlined fish exceeds that of the equivalent ‘stretched-straight’
body (see Blake, 1983a; for a discussion). Anderson et al. (2001) employed
digital particle tracking velocimetry (DPTV) to show that in still the boundary
layer flow over both scup Stenotomus chrysops L. and smooth dogfish and
Mustelus canis Mitchell is laminar at all but their highest speeds and that
swimming fish experienced greater skin friction drag than the rigid body.
Given these results it is highly likely that the swimming drag of thunniform
fish such as tuna is low. Swimming kinematics in thunniform fish are character-
ized by cyclically repeating motions and small linear and angular accelerations
of the body. Steady swimming may be sustained (patrolling, searching, migra-
tions) or prolonged (chasing).

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1202 R. W. BLAKE

Specializations in body form and kinematics associated with minimizing drag


and maximizing thrust for efficient cruising are underlain by distinct osteo-
logical and myological adaptations in tuna (Firestine & Walters, 1968). Osteo-
logical adaptations are associated with maximizing body rigidity in all but parts
of the peduncular region. In the caudal region, lateral flanges on the peduncular
vertebrae hold the region rigid in the horizontal plane and depressed neural and
heamal spines eliminate movement in the vertical plane. In addition, the caudal
fin complex is also rigid. In the post abdominal and anterior caudal regions the
neural and heamal zygapophyses are long and constrain horizontal motion
there. Post cranially, the first centrum is attached to the skull. Essentially,
flexibility is confined to a region anterior to the caudal peduncle (pre-
peduncular joint) and a region just posterior to it (post-peduncular joint).
This arrangement is similar to the dual joint system of cetaceans (Parry, 1949).
Myological specializations in tuna are associated with transferring anteriorly
produced muscle force posteriorly. Within any contracting myomere the
anterior cones exert tension on posterior vertebrae and successive posterior
myomeres are increasingly more elongate and act upon more and more
inter-vertebral joints. Force is transmitted from the bulk of the propulsive
musculature to the caudal fin by means of lateral tendons which cross the pre
and post peduncular joints.
The morphological and kinematic adaptations described above have evolved
independently in a number of phylogenetically distant groups; thunniform fish
(Scomdridae), certain sharks (Lamnidae), cetaceans and the extinct reptilian
parvipelvian ichthyosaurs (Lighthill, 1969). This broad focus on the thunniform
design is a striking example of convergent evolution. The extent of functional
evolutionary convergence is most apparent between the lamnid sharks and
tunas; in addition to a similar body form and swimming kinematics, both
groups have evolved similar myotendonous architecture, medially concentrated
red muscle that transmits force to more posterior regions of the body and caudal
fin via a system of tendons (Donley et al., 2004). Post-Triassic parvipelvian
ichthyosaurs (e.g. Stenopterygius) show some of the adaptations of tunas and
lamnid sharks for thunniform swimming [e.g. streamlined rigid body, lunate
caudal fin, similar vertebral form and presacral vertebral count, Montani
(1996)] and probably swam at optimal cruising speeds with similar metabolic
rates to tunas (Montani, 2002). Whilst later (post-Triassic) ichthyosaurs, like
Stenopterygius show evolutionary convergence with carcharhinid sharks, early
forms (e.g. Chensaurus, early Triassic) are characterized by a small caudal fin,
long narrow body, a high presacral vertebral count and probably swam in the
anguilliform mode, much like scyliorhnid sharks (Montani et al., 1996). However,
there are variations on the theme (Webb, 1984a). For example, some specialized
BCF periodic sharks (Carcharhinidae, Sphyrnidae) retain anguilliform
motions yet are characterized by a broadly thunniform design.
Recently, the long held view (based on hydromechanical theory and the
unique physiology of tunas) that the thunniform mode of propulsion is the
most energetically efficient means of BCF periodic swimming has been ques-
tioned based on comparative experimental studies of swimming performance,
kinematics and energetics in juvenile Kawakawa tuna and chub mackerel (Donley
& Dickson, 2000; Sepulveda & Dickson, 2000). The swimming performance of

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1203

juvenile ‘size matched’ E. affinis [thunniform mode, heterothermic (facultatively


endothermic)] and S. japonicus (carangiform, non-heterothermic) were
compared. From hydromechanical theory and physiological considerations it
was hypothesized that E. affinis would be a more efficient swimmer than
S. japonicus. At any given speed, E. affinis had a higher tail beat frequency,
but lower stride length, tail beat amplitude, and propulsive wavelength than
S. japonicus (Donley & Dickson, 2000). The maximum sustainable speed and
the net cost of transport (active minus resting metabolic rate), however, were
not significantly different between the two species (Sepulveda & Dickson, 2000).
Donley & Dickson (2000) and Sepulveda & Dickson (2000) concluded that the
thunniform swimmer (E. affinis) is not a more efficient swimmer than the
carangiform S. japonicus.
Explanation of these results in the context of the prevailing paradigm of the
functional design and movement pattern of fish is challenging. Sepulveda &
Dickson (2000) suggest that their results may be specific to juvenile Kawakawa
tuna and chub mackerel and imply that at larger sizes the swimming perform-
ance of heterothermic E. affinis may exceed that of the non-heterothermic
S. japonicus. The specimens of E. affinis and S. japonicus studied by Sepulveda
& Dickson (2000) had red muscle temperatures elevated relative to the water by
only 1–23 C and 01–06 C respectively. Larger tuna have an increased cap-
acity for heat exchange because rete length, rete width, and the amount of red
muscle increases with body size (Dickson et al., 2000). Dickson (1994) showed
that for five species of tuna, ranging in fork length (LF) from 155 to 667 mm
that tuna c. >210 mm are heterothermic and that the muscle temperature of
juvenile tunas are much higher than in non-heterothermic fish of the same or
greater size. Lindsey (1968) found that the red muscle temperature of a 485 mm
specimen of Kawakawa tuna E. affinis was elevated by 54 C relative to the
ocean temperature. If at higher body temperatures the specific power output of
Kawakawa tuna E. affinis exceeds that of comparably sized chub mackerel,
Sepulveda & Dickson’s (2000) suggestion may be correct. Altringham & Block
(1996), however, showed that the red muscle of non-heterothermic bonito Sarda
chiliensis Cuvier performs at least as well as that of heterothermic yellowfin tuna
at the same frequencies and temperatures. It therefore seems unlikely that adult
E. affinis will out-perform adult S. japonicus based on muscle powering.
In addition to the morphological and kinematic specializations for cruising,
tuna have distinctive anatomical and physiological adaptations of their propul-
sive musculature. The red muscle (slow contracting, aerobic) fibres of tuna
extend from a subcutaneous position to the vertebral column, as opposed to
the lateral placement typical of subcarangiform swimmers. Katz et al. (2001)
analysed red muscle function in fish based on a simple mechanical model that
treats the fish as a homogenous bending beam where muscle strain (relative
shortening) is given by the product of body curvature and the distance of the
muscle fibres from the backbone. The work done is the product of the strain
and force produced. This model argues that the internally located red muscle in
tuna should be limited in its work capacity relative to the more superficially
located red fibres. Katz et al. (2001) used sonomicrometry implanted
in yellowfin tuna to show that although the superficial red muscles conform
to the beam theory predictions as far as strain amplitude and phase relative to

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1204 R. W. BLAKE

body curvature are concerned, strain values were double those predicted and
lagged behind body curvature. Evidently, the phase difference produces brief
periods when the superficial red muscles are lengthening while the deep red
musculature is shortening and vice versa, generating a shear that doubles the net
strain and work done by the deep red muscles. In addition, the deep red muscles
at the mid-body position shorten in phase with body bending c. 20% more
posteriorly, enhancing force transmission to the caudal tendon system. Donley
et al. (2004) found a similar situation in the lamnid, shortfin mako shark (Isurus
oxyrinchus Rafinesque) and conclude that the mechanism enables both tunas
and lamnid sharks (convergent evolution) to swim in the relatively stiff-bodied
thunniform mode. The carangiform mackerels and bonitos, however, lack the
specialized red muscle adaptations of tunas (Altringham & Shadwick, 2001).
Given that there is no significant difference in the maximum sustained
swimming speed and cost of transport of size matched carangiform S. japonicus
and thunniform E. affinis (Sepulveda & Dickson, 2000) it seems reasonable to
argue that the specialized red muscle arrangements in tunas, like heterothermy,
does not translate to superior swimming performance relative to certain
carangiform swimmers.
Arguably, given roughly similar muscle power production capacity, the
sustained swimming performance of heterothermic (thunniform) swimmers
should exceed that of carangiform non-heterothermic forms based on hydro-
mechanical considerations alone. It is possible, however, that the qualitative
distinctions between the kinematics and body form of the carangiform and
thunniform modes stemming from the work of Breder (1926) are not sufficiently
robust to accommodate the significance of the scale of differences documented
by Sepulveda & Dickson (2000). Indeed, carangiform swimmers have been
described as ‘verging on the thunniform’ (Lindsey, 1978). It may be that whilst
the broad division of BCF periodic swimmers into four distinct modes is useful
in as much as each mode is characterized by kinematic and morphological traits
that are associated with different mean levels of hydromechanical efficiency
(anguilliform: low, subcarangiform: moderate, carangiform and thunniform:
high) the classification nevertheless reflects a continuum between whole body
undulation and localized oscillation and the distinctions between the swimming
performance of fish of extremely similar body form and kinematics can not be
easily discerned based on such categories. In addition, a continuum of body
forms and kinematics can be represented by different species of a single family.
According to the generally accepted classification all batoid fishes (electric
rays, sawfishes, guitarfishes, skates and stingrays) fall into the rajiform mode.
Rosenberger (2001), however, documents a range of kinematics in batoid fishes
from undulation to oscillation related to phylogenetic position and mode of life.
Heterothermy may have evolved for reasons other than increased red muscle
power output. Brill & Bushnell (1991) argue that the high cardio respiratory
performance of tunas is an adaptation to rapid oxygen debt repayment following
exhaustive activity rather than allowing for exceptional sustained swimming
performance. Block (1987) contends that an elevated, stable temperature for
sensitive organs (brain, eyes) allows tuna and other fish to experience rapid
temperature changes without adverse biochemical or physiological effects,
allowing their range of movement in the ocean to be increased. This is important

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1205

in the open ocean where food is scarce. Block (1987) also argues that warming
of the retina may increase the visual threshold in tuna, improving their ability to
detect prey. Block et al. (1993) have argued that heterothermy evolved to
maintain body temperature and muscle performance as tuna enter colder
(deeper) waters in search of prey. Brill et al. (2000), however, showed that the
heart rate of captive tuna subject to imposed temperature changes is not
protected from low ambient temperatures and therefore the aerobic per-
formance of the fish is not likely to remain high.
Katz (2002) views attempts to explain the evolutionary basis of so-called
‘elite’ performance in tunas as largely a parsimonious basis for hypothesizing
that heterothermy is part of an integrated high performance design and suggests
an alternative view based on the development of a high red muscle mass. Katz
(2002) employs a model of ‘effective’ v. ‘efficient’ aerobic metabolic strategies
utilized in the study of the physiological ecology of terrestrial vertebrates (Hayes
& Garland, 1995). Effective strategists maintain a higher aerobic speed (high red
muscle mass) than their prey and their competitive predators to maximize
feeding success. Conversely, efficient forms with small amount of red muscle
have a lower metabolic consumption thereby reducing foraging costs. Katz
(2002) views tuna as effective strategists and salmon and mackerel as efficient
strategists. Sepulveda & Dickson (2000), however, showed no significant differ-
ence between the cost of transport between Kawakawa tuna and chub mackerel
(effective and efficient strategists respectively sensu Katz, 2002). Nevertheless,
whereas studies of size matched juvenile fish may be sufficient for evaluating
comparative performance of forms utilizing different swimming modes a
comparison of metabolic rate v. swimming speed in adult heterothermic and
non-heterothermic fish is necessary to test Katz’s hypothesis.
It has been suggested that drag reduction mechanisms are likely to operate in
thunniform swimmers (Bone, 1975; Webb, 1975; Blake, 1983a). Such mech-
anisms may be viewed as adaptations that function to reduce the energy cost
and increase the hydromechanical performance of BCF periodic swimming.
Suggested possibilities include mucous as a drag reduction agent, maintenance
of a laminar boundary layer over the streamlined form through elimination of
roughness elements (recessed eyes and nares, fin slots, smooth skin), skin
structures associated with damping out inertial forces in the boundary layer
(dynamic viscous damping), the injection of gill effluent (fluid momentum) into
the boundary layer, streamlined keels and fairings, and the role of finlets
positioned on the dorsal and ventral surfaces of the caudal peduncle. Whilst
drag reduction mechanisms are expected to be particularly important in
reducing the resistance of active cruising fish and cetaceans most assessments
are based on indirect analogies with engineering structures (Rohr et al., 1998)
rather than direct experimental proof. An exception to this is the study of
Nauen & Lauder (2000) on the function of the finlets in the chub mackerel
S. japonicus. Relating experimental results based on flow visualization to rele-
vant hydromechanical models they found no significant contributions of the
finlets to either drag reduction or thrust production.
The food of many specialized BCF periodic swimmers is widely dispersed
in space and time, so that to exploit it efficiently the greatest volume of water
must be sampled in the minimum energy cost (Weihs & Webb, 1983). The

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1206 R. W. BLAKE

theoretical optimal cruising speed for maximum energy advantage is


relatively low at c. 1 bl s1 (Weihs, 1973b; Ware, 1975). The energy efficiency
of long range foraging movements and migrations can be improved through
the use of locomotor strategies such as intermittent, burst-and-glide
(negatively buoyant fish) and burst-and-coast (neutrally buoyant fish) swim-
ming. Hydromechanical models predict energy savings or range increases of
the order of 80% for fish utilizing intermittent swimming patterns (Weihs,
1973c). Holland et al. (1990) tracked the horizontal and vertical movements
of yellowfin and bigeye tuna Thunnus obesus Lowe off the coast of Hawaii.
Records confirm the necessary burst-and-glide movement patterns, although
the fish did not engage in intermittent swimming all of the time. Blake
(1983c) shows that the optimum body form (as indicated by fineness ratio:
body length divided by maximum diameter) for efficient steady (BCF periodic)
and intermittent swimming are about the same (c. 5 : 1). It follows that the
functional design requirements for efficient BCF periodic and BCF intermittent
swimming are similar and that the selective pressures that improve long range
swimming ability will enhance both steady and intermittent locomotor per-
formance. Blake (2000) gives a detailed review of intermittent propulsion in
aquatic vertebrates.
Schooling has been proposed as a means of reducing the cost of locomo-
tion in fish (Breder, 1967; Weihs, 1974). The hydromechanical model of
schooling is based on fish utilizing the fluid energy produced by their neigh-
bours and requires precise relative positioning based on a diamond pattern.
Magnuson (1978) points out that tuna do not swim in such a fashion and
therefore are not likely to see significant energy savings. Liao et al. (2003a,
b), however, showed that trout favourably position themselves and gain
energetic advantage from the vortices shed by stationary objects in their
environment (Karman gait, see p. 25). They suggest that fish may also
exploit the thrusting wakes of their neighbours when schooling. The
positional requirements for this are likely less rigorous than those required
by the ‘Weihs model’ and if so, tuna may gain hydrodynamic advantage
based on the mechanism proposed by Liao et al. (2003a, b). Tuna schooling
also may involve co-operative group behaviours in the context of feeding.
Webb & de Buffrenil (1990) point out that smaller thunniform fish (c. <05 m
in length), such as yellowfin tuna, feed on elusive prey (fish and cephalopods)
where co-operative group behaviours such as concentrating, disturbing and
disorienting prey could function to reduce the relative maneuverability of
predator and prey. Partridge et al. (1983) suggest that the group behaviour of
bluefin tuna Thunnus thynnus L. reduces the search areas of individual fish
and allows them to surround their prey. Batty & Domenici (2000) review the
kinematics and behaviour of fish and other aquatic vertebrates in the context
of predator-prey relationships.
The focus thus far has been on swimming performance and functional design
in thunniform fish in the context of other pertinent areas of tuna biology:
metabolism and energetics, anatomical and physiological specializations for
heterothermy and muscle physiology. These related fields are thoroughly
reviewed by Korsmeyer & Dewar (2001), Graham & Dickson (2001) and
Altringham & Shadwick (2001) respectively.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1207

FUNCTIONAL DESIGN, PERFORMANCE AND FEEDING IN


MEDIAN AND PAIRED FIN SWIMMERS
Median and paired fin (MPF) propulsion is common in fish that inhabit
complex environments (e.g. coral reefs, kelp beds, weedy rivers and ponds). In
addition, many BCF periodic swimmers and BCF transient specialists employ
fin-based mechanisms of propulsion for low-speed swimming and hovering. The
functional design of MPF specialists reflects the importance of different perform-
ance characteristics than those central to the predator-prey interactions of most
fusiform BCF swimmers. In particular, precise maneuver is a feature of MPF
specialists. Many undulatory and oscillatory fin swimmers are capable of turn-
ing on their own axis with little or no lateral translation of the body (Blake,
1976, 1977, 1978; in seahorse, boxfish and triggerfish respectively). However,
the traditional view that MPF forms are more maneuverable than fusiform fish
has recently been questioned. Walker (2000) defines the minimum space
required to turn as an alternate measure of maneuverability. Using this meas-
ure, rigid bodied forms (e.g. Ostraciidae, Sygnathidae) compare poorly to
flexible forms. Schrank et al. (1999) show that goldfish Carassius auratus L.
are more adept at turning around bends than either silver dollars Metynnis
hypsauchen Muller and Troschel or angelfish P. eimekei and conclude that
greater body depth and anterior-lateral pectoral fin positions are not associated
with greater maneuverability. Gerstner (1999) assessed the maneuverability
during feeding and agonistic interactions in four species of coral reef fish differing
in body and pectoral fin morphology. She found that morphological differences
among the fish were correlated with behavioural rather than perform-
ance differences. While certain deep-bodied fish with an acanthopterygian fin
distribution may not be characterized by superior maneuverability relative to
certain fusiform fish, however, many MPF fish propelled by undulatory fins
can rotate about their own axis and reverse fin waveform direction allowing
them to back out of crevices and away from obstacles (Blake, 1983d). Most
fusiform fish move backwards in a clumsy fashion based on sculling motions of
the pectoral fins. D’Aout & Aerts (1999) compared forward swimming in eels
Anguilla anguilla L. and found backwards swimming to be less hydromechanic-
ally efficient than forward motion. Backward swimming performance in many
MPF swimmers propelled by undulatory fins is as effective and efficient as
forward propulsion. In addition to releasing many MPF forms from the
morphological constraints associated with minimizing resistance (streamlining)
low absolute swimming speeds facilitates precise maneuverability.
Webb (1984a) suggests that the optimal design for MPF propulsion should
include mid-lateral pectoral fin insertion, ventro-lateral insertion of the pelvic
fins, soft-rayed dorso-ventrally symmetrical dorsal and anal fins, leading edge
reinforcement of fins, and a short, deep, laterally compressed body. There are
many examples that show one or more of these characters. The marine butter-
flyfish genus, Chaetodon, however, represents an integration of them all and
Webb (1984a) suggests the term Chaetodontiform to describe the locomotor
mode of such fish. Because of the lack of design constraints imposed by
streamlining (large protruding mobile eyes, protective flanges over the nares,
and spines are common) and the importance of non-locomotor selective

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1208 R. W. BLAKE

pressures (e.g. crypsis), however, MPF swimmers are highly variable in form
and it is unlikely that there is a single optimal form. This is consistent with the
fact that an optimal design for MPF propulsion has not been experimentally
proven to date.
Many undulatory MPF swimmers benefit from a hydromechanical inter-
action between their body and fins (Blake, 1983d; Lighthill & Blake, 1990).
Lighthill & Blake (1990) employed a type of elongated-body theory appropriate
to the analysis of elongated fishes with elongate median fins where the body
remains rigid (e.g. balistiform and gymnotiform modes; triggerfishes and weakly
electric-eels respectively). They showed that such forms experience a three-fold
momentum enhancement of useful thrust from their propulsive fins relative to
that of the fins ‘on their own’. In addition, side forces are also reduced,
minimizing energy wasting yawing motions and body drag. I suspect when
this theory is applied to other MPF forms energy savings of the same order
will be found. Blake (1983a, d) has shown that MPF undulatory swimmers and
MPF oscillatory forms are more hydromechanically efficient at low-speeds than
BCF periodic swimmers. In contrast to BCF periodic and transient specialists
where adaptations (both kinematic and morphological) for efficient propulsion
constrain forms to a few optimal designs, the principal energetic advantages of
MPF swimming (high hydromechanical efficiency and thrust enhancement and
drag reduction due to fin/body interactions) are common to many different
body forms.
Given the small intrinsic muscle mass, high hydromechanical efficiency and
thrust enhancement of the fins and low rigid body drag of many laterally
compressed MPF swimmers a low cost of swimming relative to swimming in
the BCF mode should be reflected in lower metabolic costs of transport for
MPF propulsion. Korsmeyer et al. (2002) measured oxygen consumption rates
with swimming speed in parrotfish (Scarus schlegeli, Bleeker, labriform mode)
and triggerfish [Rhinecanthus aculeatus (L.), balistiform mode] and found that
MPF swimming is energetically less costly that BCF periodic propulsion at
comparable speeds. Gordon et al. (2000) demonstrated that despite a high
body drag coefficient, the total cost of transport curves (from oxygen consump-
tion rates) for boxfish (Ostracion meleagris, Shaw and Nodder, ostraciiform
mode) is comparable to BCF periodic swimmers (including sockeye salmon) at
similar speeds. They explained this result by speculating that the negative effect
of a high body drag coefficient is approximately balanced out by lower
metabolic costs associated with reduced recoil forces in MPF propulsion
compared to BCF swimmers. However, the energy costs associated with recoil
forces in the BCF swimmers involved in the comparison are likely to be small
too. Given the probable positive metabolic implications of the locomotor
adaptations of MPF swimmers outlined above, relative recoil energy costs
may be insufficient and unnecessary to explain the similar cost of transport of
MPF and BCF swimmers.
In comparison to undulatory MPF swimmers, little work has been done on
MPF oscillatory (drag and lift based swimmers). Blake (1979a, b, 1980a, 1981a, b,
1983a, b) discusses the mechanics, energetics, and the influence of fin shape on
the propulsion of drag-based (rowing) pectoral fin (labriform) swimmers.
Walker & Westneat (1997) investigate lift-based labriform locomotion in the

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1209

bird wrasse Gomphosus varius Lacépède and point to mechanical analogies to


animal flight as far as potentially unsteady hydrodynamic effects are con-
cerned. Drucker & Jensen (1997) conducted a comparative study of kinematic
and electromyographic aspects of lift-based pectoral fin swimming in surfperches
(Embiotocidae). They found that despite of phylogenetic and ecological diver-
gence in the family mechanistic differences associated with propulsion are
small.
In contrast to the Embiotocidae, species in the family Labridae utilize a
variety of propulsive mechanisms from drag-based to lift-based pectoral fin
propulsion (Walker & Westneat, 1997). Surprisingly, a simple correlation
between pectoral fin aspect ratio and sustained swimming speed (Walker &
Westneat, 1997) is sufficient to allow for a correlation of fin morphology (aspect
ratio) and habitat use (Bellwood & Wainwright, 2001; Fulton et al., 2001).
Fulton et al. (2001) observed a distinct dichotomy between shallow high wave
action) and deep (low wave energy) reef zones off the coast of Lizard Island,
Great Barrier Reef, with high aspect ratio (faster, lift-based swimmers) and low
aspect ratio slower, drag-based forms) in greater abundance in the shallows and
deeper waters respectively. Differences in water column use were also noted, with
deeper water species tending to occur neat the substratum, exploiting refuges and
boundary layer effects (Fulton et al., 2001). Bellwood & Wainwright
(2001) observed similar patterns of distribution and water column use in
labrids across the Great Barrier Reef. If locomotor performance is a general
force shaping habitat use in coral reef fish at the familial level, other fish
propelled in other modes might be expected to show the same or similar determin-
ants of distribution as labrids. For example, the triggerfishes (Balistidae)
are propelled in the MPF undulatory mode and can be broadly divided into
Type I (high aspect ratio, high amplitude, long wavelength, low frequency) and
Type II (low aspect ratio, low amplitude, short wavelength, high frequency)
groups. These kinematic differences likely translate to significant differences in
sustained swimming performance and might constrain triggerfish species to
particular reef habitats.

MORPHOLOGICAL SPECIALIZATION, GAITS AND PERFORMANCE


TRADE-OFFS
Webb (1984a, 1998) hypothesized that specialization in any one class of
swimming (BCF transient, BCF periodic, MPF) would limit performance in
any other. This hypothesis holds for fish that employ the same propulsor for
more than one type of swimming. For example, Harper & Blake (1991) showed
that pike (BCF transient specialist) have superior fast-start performance to
rainbow trout (BCF generalist) and Webb (1988) found that the steady swim-
ming of rainbow trout is superior to that of pike. In fish that employ different
propulsive structures for different types of swimming (e.g. MPF motions for
low-speed routine swimming and BCF motions for fast-starts), however, per-
formance trade-offs do not occur. For example, Domenici & Blake (1991,
1993a, b) investigated the escape (fast-start, BCF motions) performance of
angelfish P. eimekei, a disc-shaped fish specialized for pectoral fin (labriform)
swimming at low-speeds (Blake, 1979a, b, 1980b). Comparisons of mean and

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1210 R. W. BLAKE

maximum acceleration rates with other fish (Domenici & Blake, 1991, Table. 4)
show that angelfish have good fast-start performance despite specializations for
MPF routine swimming. Kasapi et al. (1993) documented the kinematics and
performance of the BCF escape responses of the knifefish Xenomystus nigri
(Gunther). In low-speed swimming the knifefish is propelled by an undulatory
anal fin (gymnotiform mode). However, knifefish achieve high maximum
acceleration rates relative to other fish (Kasapi et al., 1993).

COUPLED AND DECOUPLED LOCOMOTOR SYSTEMS


The terms ‘coupled’ and ‘decoupled’ may be to describe systems where the
same and different locomotor structures are employed respectively. Many
specialized MPF swimmers fall into the decoupled category. Most forms char-
acterized by coupled locomotor systems are generalists (sensu Webb, 1984a).
Tuna and other thunniform swimmers are notable exceptions; adaptations (in
body form, kinematics, osteology, myology) for efficient BCF periodic
swimming and poor BCF transient (turns, accelerations) relative to many
other fish has already been noted above. The extent of specialization for
optimal steady swimming (cruising) performance necessarily implies consequent
compromises in unsteady swimming performance because the propulsive
mechanisms are coupled. Functional compromises occur at the level of force
generation (propulsive red and white muscle), transmission (force transfer to the
lateral peduncular tendons) and delivery (caudal fin). The anatomical arrange-
ments that facilitate the transmission of force produced by the red muscle fibres
to the caudal fin (Katz et al., 2001; Donley et al., 2004) allows for a functional
division between the sites of force generation and delivery within a stiff body,
allowing for near minimum rigid body drag (relative to equivalent technical
bodies of the same fineness ratio) that is significantly lower than that experi-
enced by subcarangiform swimmers.
Tuna exhibit few of the adaptations associated with good unsteady swimming
performance (i.e. large depth of section enhanced by collapsible median fins,
deep caudal peduncle, large low aspect ratio caudal fin, high degree of lateral
body flexibility, high proportion of white relative to white propulsive muscle).
Indeed, the very adaptations for optimal cruising performance in tuna comprom-
ise their unsteady swimming ability. In particular, turning and sudden accel-
erations are compromised by a lack of median fin development, a bulky oblate
spheroidal cross-section over most of the body length, a rigid vertebral column,
and a relatively low proportion of white to red muscle volume compared to
most subcarangiform fish. Tuna myomeres consist of both red and white
muscle. For unsteady propulsion, contraction of the white fibres serves to
increase the ‘pull’ on the lateral peduncular tendons and the force must be
transmitted posteriorly, such that sudden accelerations and turns are powered
solely by the caudal fin (Firestine & Walters, 1968). In addition, anterior
rigidity limits localized anterior bending and head deflection into the turn,
which in turn limits the effectiveness of directing propulsive force into the
turning direction. Presumably, directional changes in tuna are initiated by
asymmetrical tail thrusts and/or unbalanced side forces produced by the pec-
toral fins. The limitations of the coupled locomotor system of thunniform

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1211

swimmers in performing multiple tasks are mitigated by specializations for


efficient cruising, which allow large areas of open ocean to be exploited. In
addition, decrements in prey capture performance are likely compensated for
by organized and/or incidental co-operative group behaviours such as concen-
trating, disturbing, and disorienting prey, which effectively reduce the relative
maneuverability of the predator and prey (Howland, 1974; Partridge et al.,
1983). Generalist, BCF subcarangiform swimmers may be viewed as a com-
promise solution, allowing for sufficient levels of performance in both steady
and unsteady swimming in the context of coupled locomotor systems. The
compromises in body form and kinematics arise because the advantage of
near equivalent rigid body drag in cruising (realized in thunniform swimmers
through caudal transmission of propulsive muscle power) can not be achieved
without compromising the local white muscle contraction anteriorly that pro-
duces the bodily deflections necessary for effective turning and fast-starts.
Arguably, much of the adaptive radiation of teleosts reflects the evolution of
uncoupled propulsion systems that allow for multiple optimal performance
levels in more than one swimming behaviour.
Webb (1998) recast his original construct of mutually exclusive specialist
designs (BCF steady, BCF transient, MPF swimmers, Webb, 1984a) in the
context of gaits and introduces the concept of locomotor multitasking in fish.
Specialists (of all types) are characterized by specializations for a particular gait,
with consequent reductions of performance in other gaits. This corresponds to
Webb’s (1984a) definition of specialist classes and is essentially synonymous,
and consistent with the functional interpretation of coupled locomotor systems
presented here. Webb (1998) defines ‘multitasking’ fish as forms that express a
wide variety of gaits with the ability to modify their body and fin profiles to
allow for performance (to the extent possible given functional compromises) in
different gaits and states that ‘such fish may be thought of as generalists’. This
interpretation is consistent with Webb’s (1984a) view of generalist forms. Multi-
tasking, however, is not synonymous with the structural and functional basis of
decoupled locomotor systems presented here. Multitasking (sensu Webb, 1998)
involves fish employing (albeit in the context of different gaits), the same
structures (fins) for more than one task with consequent performance trade-
offs (coupled locomotor systems) as opposed to decoupled locomotor systems
(different structures for different functions) that release many fish from func-
tional design and performance compromises.
Swimming modes, categories and styles can broadly be view as gaits. The
concept of gait has been extensively applied to the study of terrestrial locomo-
tion (Alexander, 1989) and to a lesser extent to flight (Rayner, 1981). As far as I
am aware, the gait concept was first applied to fish swimming by Blake (1979a,
pp. 138–139) who gives a description of gait transitions from hovering to
forward rectilinear swimming in the blue gourami Trichogaster tricopterus (Pallas).
In hovering the pectoral fins move in antiphase in a sculling motion and the
posterior third of the dorsal fin may also be active (gait one). During slow
forward progression the fish is propelled entirely by the synchronous rowing
motions of the pectoral fins (labrifrom mode, gait two). As speed increases the
pectoral fins are supplemented by the caudal fin with the upper and lower lobes
of the fin moving from side to side in antiphase (gait three). With further

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1212 R. W. BLAKE

increases in speed the pectoral fins are folded against the sides of the body and
the fish swims in the subcarangiform mode with the caudal fin acting as a single
unit (gait four). No functional design compromises are required for any of the
propulsive structures involved in any of the four gaits (decoupled). The same
structures (pectoral fins) operate on the same mechanical principle (rowing) in
gaits one and two and this is also the case for the caudal fin (lifting mechanism)
in gaits three and four. Hove et al. (2001) documented three gaits associated
with increasing forward speed for the boxfish O. meleagris. Low-speeds are
powered by MPF asynchronous fin beats (gait one), as speed increases fin
beats become synchronous and the path of the fish more stable (gait two),
higher speeds are characterized by caudal fin propulsion in a burst-and-coast
style (gait three). This also constitutes a decoupled system. The mechanical
function of the pectoral and median fins (but not the directions in which forces
are produced) is fundamentally the same in gaits one and two (undulatory
MPF). Caudal fin swimming in gait three is not functionally compromised by
MPF propulsion in gaits one and two. In contrast, angelfish P. eimekei exhibit
three gaits (BCF transient, MPF rowing, and MPF undulatory hovering), two
coupled (MPF rowing and MPF undulatory hovering) and one decoupled (BCF
transient) (Blake, 1980b). A narrow fin base is required for good rowing
performance because ‘inboard’ portions of the fins contribute little, or negative
thrust (Blake, 1979a, b). Hovering, however, is facilitated by a broad fin base,
allowing for the necessary phase difference between independent fin rays to
produce an undulatory fin motion. The design of the pectoral fins of angelfish
reflects a design trade-off between the competing mechanical demands of row-
ing and hovering.
The functional significance of different gaits involving both decoupled and
coupled propulsive systems is reflected in different values of propulsive
efficiency. Blake (1979a, b, 1980b) shows that pectoral fin rowing is more
efficient than BCF periodic swimming at low forward speeds and suggests
that many subcarangiform swimmers switch from BCF swimming to pectoral
fin rowing at low-speeds as a result of this. However, not all gait transitions can
be explained on the basis of relative hydromechanical efficiency criteria. For
example, that BCF transient swimming hydromechanical efficiency is low rela-
tive to other gaits (Frith & Blake, 1991) is of no consequence if the object is to
escape from an attacking predator. BCF periodic swimming is more hydro-
mechanically efficient than burst-and-coast swimming (McCutcheon, 1977) but
it is also more energetically expensive (Blake, 1983c, 1991; Frith & Blake, 1991).
In addition to changing gait as a function of velocity, many fish exhibit
distinct gaits in response to flow conditions generated by stationary objects in
their environment. Drafting simply involves fish swimming in an area of
reduced flow. Although a distinct behaviour, drafting is not necessarily char-
acterized by a distinct kinematic gait. In contrast, in the presence of objects that
shed a drag wake consisting of alternating counter-rotating vortices (Kárman
vortex street), many fish synchronize (tune) their tail beat frequency and body
kinematics to that of the vortices shed by the object (Liao et al., 2003a, b). Liao
et al. (2003a, b) term this form of swimming the ‘Kárman gait’. Flow visual-
ization and electromyography on trout showed that only the anterior red axial
musculature is active and functions to allow the fish to favourably position itself

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1213

relative to the vortices (slaloming between vortices as opposed to swimming


through them), allowing for the production of propulsive (lifting) forces (Liao
et al., 2003b). Power input in the Kárman gait is minimized relative to conven-
tional BCF periodic swimming due to low muscle powering and the exploitation
of vortical energies.
The evolutionary selective pressures underlying the existence of different gaits
may be diverse. Webb (1994) argues that the number of gaits expressed within
lineages and locomotor performance range [which may be viewed from the
standpoint of gait fractionation, Webb & Gerster (2000)] has generally increased
over evolutionary time.

FISH FUNCTIONAL DESIGN, SWIMMING PERFORMANCE AND


AUTONOMOUS UNDERWATER VEHICLES (AUVs)
For many years biologists have drawn from engineering and the applied math-
ematical sciences to describe and evaluate the functional deign and perform-
ance of organisms, often in an ecological and evolutionary context. Recently,
engineers have become interested in the physio-mechanical designs evolved in
fish as a basis for constructing robotic autonomous underwater vehicles
(AUVs). Biomimetic designs based on BCF steady and transient swimming
and undulatory and oscillatory MPF propulsion are viewed as having great
potential for exploitation in artificial systems (Sfakiotakis et al., 1999). Argu-
ably, many of the ideas, concepts, theoretical and experimental perspectives
discussed above should be relevant to AUV design and function. Engineering
designs for single or multiple functions, however, are not constrained by
limitations imposed by phylogenetic (historical) or ontogenetic (developmental)
factors. In addition, multiple, optimal structural and functional solutions may
be found to design problems by selecting appropriate design features from a
variety of ‘fish models’. Some recent developments and future possibilities for
AUV design and function based on the form, propulsion systems and per-
formance of fish are discussed below.
Remotely releasable instruments attached to free-ranging aquatic animals are
commonly employed in studies of physiological ecology (see Andrews, 1998a, b
for reviews). Recently interest has developed in employing fish locomotion as a
model for instrumented unmanned mini-submarines (Barret et al., 1999; Gordon
et al., 2000; Triantafyllou et al., 2000; Wakefield, 2002). These autonomous
underwater vehicles (AUVs) are driven by propellers, equipped with sensors
and used for oceanographic research (e.g. mapping the three-dimensional
characteristics of the water column), industrial assessment (e.g. determining the
concentrations of pollutants) and military applications. Broadly, current AUVs
may be classified in three categories; simple, rigid mini-submarines propelled by a
single posterior propeller (Wakefield, 2002), complex multi-propeller designs that
mimic MPF swimmers (Gordon et al., 2000), and complex designs based on an
analogy with BCF swimmers (Triantafyllou et al., 2000).
Gordon et al. (2000) point out that boxfishes (Ostraciidae) are a promising
model for multi-propeller MPF swimming. Their study of the swimming
mechanics and energetics of O. meleagris concludes that they are relatively
fast, well-controlled, stable swimmers in both sustained and prolonged

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1214 R. W. BLAKE

swimming. In addition to this and the other hydromechanical advantages men-


tioned above, ostraciifrom fish are steered by the action of a large caudal fin,
which is collapsed during forward rectilinear swimming (presumably to reduce
drag, Blake, 1977). Bartol et al. (2003) employed digital particle imaging velo-
cimetry, pressure and force measurements to show that the bony carapace of the
boxfish Lactophrys triqueter L. generates trimming forces that self-correct for
instabilities associated with pitching and yawing motions. Hydrodynamic stabi-
lity is advantageous in the high energy, variable velocity environments (coral
reef shallows subject to wave action) that the fish inhabits (Bartol et al., 2003).
The carapace performs effectively in generating stabilizing moments based on
lifting forces. Lift coefficients for the carapace are comparable to those of delta
wings. Boxfish, however, are characterized by high drag coefficients (Blake,
1983b; Bartol et al., 2003) and a high acceleration reaction (inertial response
associated with the fish mass and the water that it entrains as it moves unstead-
ily). These forces will cause displacement of the fish in surging waters, albeit in a
stable fashion. This may have important behavioural and ecological con-
sequences, particularly in regard to the capacity of the fish to remain on feeding
sites and to establish and defend territory. By extension, this may have implica-
tions for AUV designs based on MPF boxfish models, especially if the AUV is
required to perform localized, site-specific tasks. A biomimetic approach to
AUV design incorporating these features of ostraciiform fish with a streamlined
body seems promising. Indeed, Gordon et al. (2000) view boxfish as ‘naturally
evolved vertebrate autonomous underwater vehicles’.
Triantafyllou & Triantafyllou (1995) and Barret et al. (1999) have explored
thunniform BCF periodic propulsion in tunas as a biomimetic model for AUVs.
A working robotic model (static test platform) mimicing bluefin tuna (‘Robotuna’)
demonstrated drag reduction (less than that predicted for an oscillating body at the
experimental Reynolds numbers of c. 106) of the order of 50%. Laminarization
of the boundary layer by the motions of the body and favourable flow
interactions between the body and fins were offered as possible explanations
for this important result. Anderson & Chhabra (2002) used the optimum body
form and tail kinematics of Robotuna (Triantafyllou & Triantafyllou, 1995) in
a free swimming AUV based on yellow fin tuna focusing on maneuverability
and stability performance. They noted that whilst tuna are not highly
maneuverable compared to many fish (Blake et al., 1995), their tuna inspired
AUV (Draper Laboratory Vorticity Control Undersea Vehicle) out performed
other conventional rigid underwater vehicles with an average turning diameter
similar to that determined by Blake et al. (1995) for T. albacares (c. 1 bl s1).
Bandyopadhyay (2002) plots a coefficient of normal acceleration v. normalized
turning radius (turning radius divided by body length) for carangiform fish and
two conventional (rigid cigar shaped) AUVs showing that the fish can make
the same radius turn as the AUVs at a normal acceleration that is lower by a
factor of c. 10.
Triantafyllou et al. (2000) focus on a BCF unsteady model for AUVs. BCF
unsteady motions allow fish to generate large forces of short duration efficiently
and co-ordinate transient motions of the body such that the energy lost in the
wake is minimized during maneuvers (Ahlborn et al., 1991, 1997) suggesting
that certain AUV designs could incorporate aspects of BCF unsteady swimming

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1215

characteristics. In addition, the ‘thrust reversal’ mechanism of swimming


described by Ahlborn et al. (1991) provides a basis for understanding how
this could be done with a minimum of hydrodynamic noise through the elim-
ination of the wake signature as the fish (or possibly an AUV) propels itself by
successively generating and destroying tail tip vortices of alternating direction.
Bandyopadhyay (2002) proposes that stealth issues in AUV design may be
addressed by reducing the rotational propulsor speed combined with a reduc-
tion of hydrodynamic noise through a means of active elimination by a rotor
blade traversing the wake. However, as Bandyopadhay (2002) acknowledges,
these potential solutions pose considerable practical engineering challenges, and
as implied above though sufficient they may not be necessary.
Anderson & Chhabra (2002) point out that AUV missions require a variety of
capabilities which may be mutually exclusive (e.g. high forward speed, long
range, maneuverability, station holding) and that ‘engineers have the luxury to
select the most desirable attributes to mimic in their designs’. However, many
solutions requiring mutually exclusive performance attributes may lie in fish
that employ decoupled gaits in the context of multiple tasks and on under-
standing the mechanics of postural control and swimming trajectories of fishes.
Weihs (2002) reviews stability criteria in relation to maneuverability in aquatic
locomotion. Webb (2002) considers stability criteria in the context of the
requirements and capacity of fish to self-correct through damping and trimming
mechanisms that stabilize swimming trajectories.
It is likely that rather than any particular fish form and swimming mode
forming the basis for AUV design a variety of designs will emerge based on
appropriate functional analogies reflecting the particular purpose of the vehicle.
In this sense, the process of engineering AUVs may be viewed as one of
convergent evolution with nature’s solutions.

CONCLUDING REMARKS
Lighthill (1969) concluded his survey of aquatic propulsion by emphasizing
the need for further research on the carangiform and thunniform swimming
modes. Many of the issues that he raised (hydromechanical, physiological, field-
based performance measurements) have been addressed. However, among other
things, integrated studies of the hydromechanics, comparative swimming
performance and energetics, and muscle physiology of large (adult) carangiform
and thunniform fishes are needed to establish whether or not ‘optimal’ perform-
ance is solely associated with the thunniform mode with a possible role for
heterothermy. In addition, further experimental studies on the possible role in
drag reduction in thunniform fishes of structures such as keels, finlets and scale
corselets would be worthwhile.
As far as I am aware, whilst there is a firm theoretical basis for predicting
significant thrust enhancement and drag reduction due to fin/body interactions
in MPF swimmers there are no comprehensive comparative experimental stud-
ies on this issue. Determinations of the implications for swimming energetics in
relation to body shape, fin form and kinematics should be considered in a
behavioural and ecological context (activity level, swimming styles, habitat).
From an evolutionary standpoint, the phylogenetic relationships, if any, of fin

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1216 R. W. BLAKE

form and kinematics and body shape should also be assessed. In addition to
giving further insight into understanding body and fin form and kinematics in
relation to swimming energetics in MPF swimmers, studies of thrust enhance-
ment and drag reduction due to fin/body interactions could have implications
for the design and operation of AUVs based on the MPF model.
Consideration should be given to further studies of functional design and
swimming kinematics in the context of coupled and decoupled gaits. In par-
ticular, comparative studies of the relative thrust, power requirements, and
hydromechanical efficiency of different gaits in relation to swimming speed
would allow for a mechanistic understanding of gait selection and transitions
as a function of swimming speed among and between coupled and decoupled
gaits. Further studies on the mechanism and context of use of the Kárman and
related gaits would be rewarding. The experiments of Liao et al. (2003a, b)
involved vorticity shed by simple structures (cylinders bisected along their long
axis) generating one type of drag wake (Kárman vortex street) with which the
fish could interact, more complex arrangements of different geometries that
more closely represent the actual ecological niches of fish would be useful in
assessing the overall energetic significance of such gaits in nature. In addition to
drafting and the exploitation of the wake energy of objects (including other
fish), fish may also use fluid energies (detected by the inner ear and lateral line)
together with other sensory information (visual system) to facilitate object
detection and avoidance. This information could be relevant to the design and
operation of AUVs and fish management (e.g. fishway design).

The author would like to thank W. Megill and two anonymous referees for comments
on an earlier draft of this paper.

References
Ahlborn, B., Harper, D., Blake, R., Ahlborn, D. & Cam, M. (1991). Fish without
footprints. Journal of Theoretical Biology 148, 521–533.
Ahlborn, B., Chapman, S., Stafford, R., Blake, R. W. & Harper, D. G. (1997).
Experimental simulation of the thrust phases of fast-start swimming of fish.
Journal of Experimental Biology 200, 2301–2312.
Alexander, R. McN. (1989). Optimization and gaits in the locomotion of vertebrates.
Physiological Reviews 69, 1199–1227.
Altringham, J. D. & Block, B. A. (1996). Why do tuna maintain elevated slow muscle
temperatures? Power output of muscle isolated from endothermic and ectothermic
fish. Journal of Experimental Biology 200, 2617–2627.
Altringham, J. D. & Shadwick, R. E. (2001). Swimming and muscle function. In Tuna
Physiology, Ecology, and Evolution (Block, B. A. & Stevens, E. D., eds),
pp. 313–344. London: Academic Press.
Anderson, E. J. & Chhabra, N. K. (2002). Maneuvering and stability performance of a
robotic tuna. Integrative and Comparative Biology 42, 118–126.
Anderson, J. M., McGillis, W. R. & Grosenbaugh, M. A. (2001). The boundary layer of
swimming fish. Journal of Experimental Biology 204, 81–102.
Andrews, R. D. (1998a). Instrumentation for the remote monitoring of physiological and
behaviourial variables. Journal of Applied Physiology 85, 1974–1981.
Andrews, R. D. (1998b). Remotely releasable instruments for monitoring the foraging
behaviour of pinnipeds. Marine Ecological Progress Series 175, 289–294.
Bainbridge, R. (1960). Speed and stamina in three fish. Journal of Experimental Biology
37, 129–153.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1217

Bandyopadhyay, P. R. (2002). Maneuvering hydrodynamics of fish and small underwater


vehicles. Integrative and Comparative Biology 42, 102–117.
Barret, D. S., Triantafyllou, M. S., Yue, D. K. P., Grosenbaugh, M. A. & Wolfgang, M. J.
(1999). Drag reduction in fish-like locomotion. Journal of Fluid Mechanics 392,
183–212.
Bartol, I. K., Gharib, M., Weihs, D., Webb, P. W., Hove, J. R. & Gordon, M. S. (2003).
Hydrodynamic stability of swimmimg in ostraciid fishes: role of the carapace in the
smooth trunkfish Lactophrys triqueter (Teleostei: Ostraciidae). Journal of
Experimental Biology 206, 725–744.
Batty, R. S. & Domenici, P. D. (2000). Predator-prey interactions in fish and other
aquatic vertebrates: kinematics and behaviour. In Biomechanics in Animal
Behaviour (Domenici, P. D. & Blake, R. W., eds), pp. 237–257. Oxford: Bios
Scientific Publishers.
Beamish, F. W. H. (1978). Swimming capacity. In Fish Physiology, Vol. 7 (Hoar, W. S. &
Randall, D. J., eds), pp. 101–187. New York, London: Academic Press.
Behnke, R. J. (1972). The systematics of salmonid fishes of recently glaciated lakes.
Journal of the Fisheries Research Board of Canada 29, 639–671.
Bellwood, D. R. & Wainwright, P. C. (2001). Locomotion in labrid fishes: implications
for habitat use and cross-shelf biogeography on the Great Barrier Reef. Coral
Reefs 20, 139–150.
Blake, R. W. (1976). On seahorse locomotion. Journal of the Marine Biological
Association of the United Kingdom 56, 939–949.
Blake, R. W. (1977). On ostraciiform locomotion. Journal of the Marine Biological
Association of the United Kingdom 57, 1047–1055.
Blake, R. W. (1978). On balistiform locomotion. Journal of the Marine Biological
Association of the United Kingdom 58, 73–80.
Blake, R. W. (1979a). The Mechanics of Labriform Locomotion. Unpublished PhD
thesis, University of Cambridge.
Blake, R. W. (1979b). The mechanics of labriform locomotion I: Labriform locomotion in
the Angelfish (Pterophyllum eimekei): an analysis of the power stroke. Journal of
Experimental Biology 82, 255–271.
Blake, R. W. (1980a). Undulatory median fin propulsion of two species of teleost with
different modes of life. Canadian Journal of Zoology 58, 2116–2119.
Blake, R. W. (1980b). The mechanics of labriform locomotion II: An analysis of the
recovery stroke and the overall fin-beat propulsive efficiency in the Angelfish.
Journal of Experimental Biology 85, 333–342.
Blake, R. W. (1981a). Influence of pectoral fin shape on thrust and drag in labriform
locomotion. Journal of Zoology (London) 194, 53–66.
Blake, R. W. (1981b). Mechanics of drag-based mechanisms of propulsion in aquatic
vertebrates. In Vertebrate Locomotion (Day, M. H., ed.). Symposium of the
Zoological Society of London 48, 29–52. London: Academic Press.
Blake, R. W. (1983a). Fish Locomotion. Cambridge: Cambridge University Press.
Blake, R. W. (1983b). Median and paired fin propulsion. In Fish Biomechanics
(Webb, P. W. & Weihs, D., eds), pp. 214–247. New York: Praeger Press.
Blake, R. W. (1983c). Functional design and burst-and-coast swimming in fishes.
Canadian Journal of Zoology 61, 2491–2494.
Blake, R. W. (1983d). Swimming in the electric-eels and knifefishes. Canadian Journal of
Zoology 61, 1432–1441.
Blake, R. W. (1991). On the efficiency of energy transformations in cells and animals. In
Efficiency and Economy in Animal Physiology (Blake, R. W., ed.), pp. 13–32.
Cambridge: Cambridge University Press.
Blake, R. W. (2000). The biomechanics of intermittent swimming behaviours in aquatic
vertebrates. In Biomechanics in Animal Behaviour (Domenici, P. D. & Blake, R. W.,
eds), pp. 79–100. Oxford: Bios Scientific Publishers.
Blake, R. W., Chatters, L. M. & Domenici, P. D. (1995). Turning radius of yellowfin tuna
(Thunnus albacares) in unsteady swimming maneuvres. Journal of Fish Biology 46,
536–538.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1218 R. W. BLAKE

Blaxter, J. H. S. & Fuiman, L. A. (1990). The role of the sensory systems of herring larvae
in evading predatory fishes. Journal of the Marine Biological Association of the
United Kingdom 70, 413–427.
Block, B. A. (1987). Strategies for regulating brain and eye temperatures: a thermogenic
tissue. In Comparative Physiology: Life in Water and on Land (Dejours, P., Bolis, L.,
Taylor, C. R. & Weibel, E. E., eds), pp. 401–420. Padova: Fida Research Series IX.
Block, B. A., Finnerty, J. R., Stewart, A. F. R. & Kidd, J. (1993). Evolution of
endothermy in fish: mapping physiological traits on a molecular phylogeny.
Science 260, 210–214.
Boisclair, D. & Tancy, M. (1993). Empirical analysis of the influence of swimming pattern
on the net energetic cost of swimming fishes. Journal of Fish Biology 42, 169–183.
Bone, Q. (1975). Muscular and energetic aspects of fish swimming. In Swimming and
Flying in Nature, Vol. 2 (Wu, T. Y., Brokaw, C. J. & Brennan, C., eds),
pp. 493–528. New York: Plenum Press.
Breder, C. M. (1926). The locomotion of fishes. Zoologica 4, 159–297.
Breder, C. M. (1967). On the survival value of fish schools. Zoologica 52, 25–40.
Breinerd, E. L. & Patek, S. N. (1998). Vertebral column morphology, C-start curvature, and
the evolution of mechanical defenses in tetradontiform fishes. Copeia 1998, 971–984.
Brett, J. R. (1964). The respiratory metabolism and swimming performance of
young sockeye salmon. Journal of the Fisheries Research Board of Canada
21, 1183–1226.
Brill, R. W. & Bushnell, P. G. (1991). Metabolic and cardiac scope of high energy demand
teleosts, the tunas. Canadian Journal of Zoology 69, 2002–2009.
Brill, R. W., Lowe, T. E. & Cousins, K. L. (2000). Hot water temperature really limits the
vertical movements of tunas and billfishes – it’s the heart stupid. Pelagic Fisheries
Research Program. JIMAR-SOEST, University of Hawaii.
D’Aout, K. & Aerts, P. (1999). A kinematic comparison of forward and backward
swimming in the eel Anguilla anguilla. Journal of Experimental Biology 202, 1511–1521.
Dickson, K. A. (1994). Tunas as small as 207 mm fork length can elevate muscle
temperatures significantly above ambient water temperature. Journal of
Experimental Biology 190, 79–93.
Dickson, K. A., Johnson, N. M., Hoskinson, J. A., Hanson, M. W. & D’Souza, J. (2000).
Ontogenetic changes in characteristics required for endothermy in juvenile black
skipjack tuna (Euthynnus lineatus). Journal of Experimental Biology 203, 3077–3087.
Dill, L. M. (1974). The escape response of the zebra danio (Brachydanio rerio). I. The
stimulus for escape. Animal Behaviour 22, 710–721.
Domenici, P. D. (2001). The scaling of locomotor performance in predator-prey encounters:
from fish to killer whales. Comparative Biochemistry and Physiology A 131, 169–182.
Domenici, P. D. & Blake, R. W. (1991). The kinematics and performance of the escape
response in the angelfish (Pterophyllum eimekei). Journal of Experimental Biology
156, 187–205.
Domenici, P. D. & Blake, R. W. (1993a). The effect of size on the kinematics and
performance of angelfish (Pterophyllum eimekei). Canadian Journal of Zoology
71, 2319–2326.
Domenici, P. D. & Blake, R. W. (1993b). Escape trajectories in angelfish (Pterophyllum
eimekei). Journal of Experimental Biology 177, 253–272.
Domenici, P. D. & Blake, R. W. (1997). The kinematics and performance of fast-start
swimming. Journal of Experimental Biology 200, 1165–1178.
Domenici, P. D. & Blake, R. W. (2000). Biomechanics in behaviour. In Biomechanics in
Animal Behaviour (Domenici, P. D. & Blake, R. W., eds), pp. 1–17. Oxford: Bios
Scientific Publishers.
Donley, J. M. & Dickson, K. A. (2000). Swimming kinematics of juvenile Kawakawa
Tuna (Euthynnus affinis) and Chub Mackerel (Scomber japonicus). Journal of
Experimental Biology 203, 3103–3116.
Donley, J. M., Sepulveda, C. A., Konstantinidis, P., Gemballa, S. & Shadwick, R. E.
(2004). Convergent evolution in mechanical design of lamnid sharks and tunas.
Nature 429, 61–65.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1219

Drucker, E. G. & Jensen, J. S. (1997). Kinematic and electromyographic analysis of


steady pectoral fin swimming in the surfperches. Journal of Experimental Biology
200, 1709–1723.
Eaton, R. C., Bombardieri, R. A. & Meyer, D. L. (1977). The Mauthner-initiated startle
response in fish. Journal of Experimental Biology 66, 65–81.
Eaton, R. C. & Hackett, J. T. (1984). The role of Mauthner cell in fast-starts involving
escape in teleosts. In Neural Mechanisms of Startle Behaviour (Eaton, R. C., ed.),
pp. 213–266. New York: Plenum Press.
Ehlinger, T. J. (1990). Habitat choice and phenotype limited feeding efficiency in Bluegill:
individual differences and trophic polymorphisms. Ecology 71, 886–896.
Ehlinger, T. J. & Wilson, D. S. (1988). Complex foraging polymorhphism in bluegill
sunfish. Proceedings of the National Academy of Sciences of the USA 85, 1878–1882.
Farlinger, S. & Beamish, F. W. H. (1977). Effects of time and velocity increments on the
critical swimming speed of largemouth bass (Micropterus salmoides). Transactions
of the American Fisheries Society 106, 436–439.
Frith, H. R. & Blake, R. W. (1991). Mechanics of the startle response in northern pike,
Esox lucius. Canadian Journal of Zoology 69, 2831–2839.
Frith, H. D. & Blake, R. W. (1995). The mechanical power output and hydrodynamic
efficiency of northern pike (Esox lucius) fast-starts. Journal of Experimental
Biology 198, 1863–1873.
Fulton, C. J., Bellwood, D. R. & Wainwright, P. C. (2001). The relationship between
swimming ability and habitat use in wrasses (Labridae). Marine Biology 139, 25–33.
Gero, D. R. (1952). The hydrodynamic aspects of fish propulsion. American Museum
Notivates 1601, 1–32.
Gerstner, C. L. (1999). Maneuverability of four species of coral reef fish that differ in
body and pectoral fin morphology. Canadian Journal of Zoology 77, 1102–1110.
Gillis, G. B. (1997). Anguilliform locomotion in an elongate salamander (Siren
intermedia): effects of speed on axial undulatory movements. Journal of
Experimental Biology 200, 767–784.
Gordon, M. S., Hove, J. R., Webb, P. W. & Weihs, D. (2000). Boxfishes as unusually well
controlled autonomous underwater vehicles. Physiological and Biochemical
Zoology 73, 663–671.
Graham, J. B. & Dickson, K. A. (2001). Anatomical and physiological specialization for
endothermy. In Tuna: Physiology, Ecology, and Evolution (Block, B. A. & Stevens,
E. D., eds), pp. 121–165. London: Academic Press.
Hammer, C. (1995). Fatigue and exercise tests with fish. Comparative Biochemistry and
Physiology 112(A), 1–20.
Harper, D. G. & Blake, R. W. (1988). Energetics of piscivorous predator-prey
interactions. Journal of Theoretical Biology 134, 59–76.
Harper, D. G. & Blake, R. W. (1989a). A critical analysis of the use of high-speed film to
determine maximum acceleration in fish. Journal of Experimental Biology 142,
465–471.
Harper, D. G. & Blake, R. W. (1989b). On the error involved in high-speed film when used to
evaluate maximum accelerations of fish. Canadian Journal of Zoology 67, 1929–1936.
Harper, D. G. & Blake, R. W. (1990). Prey capture and the fast-start performance of
rainbow trout Salmo gairdneri and northern pike Esox lucius. Journal of
Experimental Biology 150, 321–342.
Harper, D. G. & Blake, R. W. (1991). Prey capture and the fast start performance of
northern pike, Esox lucius. Journal of Experimental Biology 155, 175–192.
Hayes, J. P. & Garland, T. (1995). The evolution of endothermy: testing the aerobic
capacity model. Evolution 49, 836–847.
Hertel, H. (1966). Structure, Form and Movement. New York: Reinhold.
Hove, J. R., O’Brien, L. M., Gordon, M. S., Webb, P. W. & Weihs, D. (2001). Boxfishes
(Teleostei: Ostraciidae) as a model system for fishes swimming with many fins:
kinematics. Journal of Experimental Biology 204, 1459–1471.
Howland, H. C. (1974). Optimal strategies for predator avoidance: The relative importance
of speed and maneuverability. Journal of Theoretical Biology 47, 333–350.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1220 R. W. BLAKE

Hughes, N. F. & Kelly, L. H. (1996). A hydrodynamic model for estimating the energetic
cost of swimming maneuvers from a description of their geometry and dynamics.
Canadian Journal of Fisheries and Aquatic Sciences 53, 2484–2493.
Jonsson, B. & Jonsson, N. (2001). Polymorphism and speciation in Arctic charr. Journal
of Fish Biology 58, 605–638.
Kasapi, M. A., Domenici, P. D., Blake, R. W. & Harper, D. G. (1993). The kinematics
and performance of the escape response in the knifefish, Xenomystus nigri.
Canadian Journal of Zoology 71, 189–195.
Katz, S. L. (2002). Design of heterothermic muscle in fish. Journal of Experimental
Biology 205, 2251–2266.
Katz, S. L., Syme, D. A. & Shadwick, R. E. (2001). High speed swimming enhanced
power in yellow fin tuna. Nature 410, 770–771.
Kolok, A. S. (1992). The swimming performance of individual largemouth bass (Micropterus
salmoides) are repeatable. Journal of Experimental Biology 170, 265–270.
Korsmeyer, K. E. & Dewar, H. (2001). Tuna metabolism and energetics. In Tuna:
Physiology, Ecology, and, Evolution (Block, B. A. & Stevens, E. D., eds),
pp. 35–78. London: Academic Press.
Korsmeyer, K. E., Steffensen, J. F. & Herskin, K. (2002). Energetics of median and paired
fin swimming, body and caudal fin swimming, and gait transition in parrot fish
(Scarus schlegeli) and triggerfish (Rhinecanthus aculeatus). Journal of Experimental
Biology 205, 1253–1263.
Law, T. C. & Blake, R. W. (1996). Comparison of the fast-start performance of closely
related morphologically distinct threespine sticklebacks (Gasterosteus spp.).
Journal of Experimental Biology 199, 2595–2604.
Lee, G. G., Farrell, A. P., Lotto, A., Hinch, S. G. & Healey, M. C. (2003a). Excess
post-exercise oxygen consumption in adult sockeye (Oncorhynchus nerka) and coho
(O. kisutch) salmon following critical speed swimming. Journal of Experimental
Biology 206, 3253–3260.
Lee, G. G., Farrell, A. P., Lotto, A., MacNutt, M. J., Hinch, S. G. & Healey, M. C.
(2003b). The effect of temperature on swimming performance and oxygen
consumption in adult sockeye (Oncorhynchus nerka) and coho (O. kisutch) salmon
stocks. Journal of Experimental Biology 206, 3239–3251.
Liao, C. L., Beal, D. N., Lauder, G. V. & Triantafyllou, M. S. (2003a). The Kárman gait:
novel body kinematics of rainbow trout swimming in a vortex street. Journal of
Experimental Biology 206, 1059–1073.
Liao, C. L., Beal, D. N., Lauder, G. V. & Triantafyllou, M. S. (2003b). Fish exploiting
vortices decrease muscle activity. Science 302, 1566–1569.
Lighthill, M. J. (1969). Hydromechanics of aquatic propulsion: a survey. Annual Review
of Fluid Mechanics 1, 413–446.
Lighthill, M. J. (1975). Mathematical Biofluiddynamics. Philadelphia: Society for Applied
and Industrial Mathematics.
Lighthill, M. J. & Blake, R. W. (1990). Biofluiddynamics of balistiform and gymnotiform
locomotion. Part I. Biological background and analysis by elongated body theory.
Journal of Fluid Mechanics 212, 187–207.
Lindsey, C. C. (1968). Temperatures of red and white muscle in recently caught marlin
and other large tropical fish. Journal of the Fisheries Research Board of Canada 25,
1987–1992.
Lindsey, C. C. (1978). Form, function and locomotory habits in fish. In Fish Physiology,
Vol. 7 (Hoar, W. S. & Randall, D. J., eds), pp. 1–100. London: Academic Press.
McCutcheon, C. W. (1977). Froude propulsive efficiency of a small fish measured by
wake visualization. In Scale Effects in Animal Locomotion (Pedley, T. J., ed.),
pp. 339–363. London: Academic Press.
Nauen, J. C. & Lauder, G. V. (2000). Locomotion in scombrid fishes: morphology and
kinematics of the finlets of the chum mackerel, Scomber japonicus. Journal of
Experimental Biology 203, 2247–2259.
O’Steen, S., Cullum, A. J. & Bennett, A. F. (2002). Rapid evolution of escape
performance in Trinidad guppies (Poecillia reticulatus). Evolution 56, 776–784.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
FISH DESIGN AND SWIMMING PERFORMANCE 1221

Parry, D. A. (1949). The swimming of whales and a discussion of Gray’s Paradox. Journal
of Experimental Biology 26, 24–34.
Partridge, B. L., Johansson, J. & Kalish, J. (1983). The structure of schools of giant
bluefin tuna in Cape Cod bay. Environmental Biology of Fishes 9, 253–262.
Rayner, J. M. V. (1981). Flight adaptations in vertebrates. In Vertebrate Locomotion
(Day, M. H., ed.). Symposium of the Zoological Society of London 48, 137–172.
London: Academic Press.
Rohr, J., Latz, M. I., Fallon, S., Nauen, J. C. & Hendricks, E. (1998). Experimental
approaches towards interpreting dolphin–stimulated bioluminescence. Journal of
Experimental Biology 201, 1447–1460.
Rome, L. C. (1995). Influence of temperature on muscle properties in relation to swimming
performance. In Biochemistry and Molecular Biology of Fishes, Vol. 5 (Hochachka, P. W.
& Mommsen, T. P., eds), pp. 73–99. Amsterdam: Elsevier Science.
Rosenberger, L. J. (2001). Pectoral fin locomotion in batoid fishes: undulation versus
oscillation. Journal of Experimental Biology 204, 379–394.
Schrank, A. J., Webb, P. W. & Mayberry, S. (1999). How do body and paired-fin
positions affect the ability of three fishes to maneuver around bends? Canadian
Journal of Zoology 77, 203–210.
Schriefer, J. E. & Hale, M. E. (2004). Strikes and startles of northern pike (Esox lucius): a
comparison of muscle activity and kinematics between S-start behaviors. Journal of
Experimental Biology 207, 535–544.
Sepulveda, C. & Dickson, K. A. (2000). Maximum sustainable speeds and cost of
swimming in juvenile Kawakawa Tuna (Euthynnus affinis) and Chub Mackerel
(Scomber japonicus). Journal of Experimental Biology 203, 3089–3101.
Sfakiotakis, M., Lane, D. M. & Davies, B. C. (1999). Review of fish swimming modes for
aquatic locomotion. IEEE Journal of Oceanic Engineering 24, 237–251.
Skulason, S., Noakes, L. G. & Snorrason, S. S. (1989). Ontogeny of trophic morphology
in four sympatric morphs of char Salvelinus alpinus in Thinguallauatin, Iceland.
Biological Journal of the Linnean Society 38, 281–301.
Sparenberg, J. A. (2002). Survey of the mathematical theory of fish locomotion. Journal
of Engineering Mathematics 44, 395–448.
Spierts, I. L. Y. & van Leeuwen, J. L. (1999). Kinematics and muscle dynamics of C- and
S-starts of carp (Cyprinus carpio). Journal of Experimental Biology 202, 393–406.
Taylor, E. B. & Foote, C. J. (1991). Critical swimming velocities of juvenile sockeye
salmon and kokanee, the anadromous and non-anadromous forms of
Oncorhynchus nerka (Walbaum). Journal of Fish Biology 38, 407–419.
Taylor, E. B. & McPhail, J. D. (1986). Prolonged and burst swimming in anadromous and
freshwater sticklebacks, Gasterosteus aculeatus. Canadian Journal of Zoology 64, 416–420.
Triantafyllou, M. S. & Triantafyllou, G. S. (1995). An efficient swimming machine.
Scientific American 272, 64–70.
Triantafyllou, M. S., Triantafyllou, G. S. & Yue, D. K. P. (2000). Hydrodynamics of
fish-like swimming. Annual Review of Fluid Mechanics 32, 33–53.
Videler, J. J. (1993). Fish Swimming. London: Chapman & Hall.
Wakefield, J. (2002). Mimicing mother nature. Scientific American 286, 26–27.
Wakeling, J. M. & Johnston, I. A. (1999). Body bending during fast-starts in fish can be
explained in terms of muscle torque and hydrodynamic resistance. Journal of
Experimental Biology 202, 675–682.
Wakeling, J. M., Kemp, K. M. & Johnston, I. A. (1999). The biomechanics of fast-starts
during ontogeny in the common carp, Cyprinus carpio. Journal of Experimental
Biology 202, 3057–3067.
Walker, J. A. (2000). Does a rigid body limit maneuverability? Journal of Experimental
Biology 203, 3391–3396.
Walker, J. A. & Westneat, M. W. (1997). Labriform propulsion in fishes: kinematics of
flapping aquatic flight in the bird wrasse Gomphosus varius (Labridae). Journal of
Experimental Biology 200, 1549–1569.
Ware, D. M. (1975). Growth, metabolism, and optimal swimming speed of a pelagic fish.
Journal of the Fisheries Research Board of Canada 32, 33–41.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222
1222 R. W. BLAKE

Webb, P. W. (1975). Hydrodynamics and energetics of fish propulsion. Bulletin of the


Fisheries Research Board of Canada 190, 1–159.
Webb, P. W. (1976). The effect of size on the fast-start performance of rainbow trout
Salmo gairdneri and a consideration of piscivorous predator-prey interactions.
Journal of Experimental Biology 65, 157–177.
Webb, P. W. (1977). Effects of median-fin amputation on fast-start performance of
rainbow trout (Salmo gairdneri). Journal of Experimental Biology 68, 123–135.
Webb, P. W. (1978a). Temperature effects on acceleration of rainbow trout Salmo
gairdneri. Journal of the Fisheries Research Board of Canada 35, 1417–1422.
Webb, P. W. (1978b). Fast-start performance and body form in seven species of teleost
fish. Journal of Experimental Biology 74, 211–226.
Webb, P. W. (1983). Speed, acceleration and maneuverability of two teleost fishes.
Journal of Experimental Biology 102, 115–122.
Webb, P. W. (1984a). Body form, locomotion and foraging in aquatic vertebrates.
American Zoologist 24, 107–120.
Webb, P. W. (1984b). Chase response latencies of some teleostean piscivores. Comparative
Biochemistry and Physiology 97(A), 45–48.
Webb, P. W. (1986). Effect of body form and response threshold on the vulnerability
of four species of teleost attacked by largemouth bass (Micropterus salmoides).
Canadian Journal of Fisheries and Aquatic Science 43, 763–771.
Webb, P. W. (1988). Steady swimming kinematics of tiger musky, an escociform accelerator,
and rainbow trout, a cruiser generalist. Journal of Experimental Biology 138, 51–69.
Webb, P. W. (1994). The biology of fish swimming. In Mechanics and Physiology of
Animal Swimming (Maddock, L., Bone, Q. & Rayner, J. M. V., eds), pp. 45–62.
Cambridge: Cambridge University Press.
Webb, P. W. (1998). Swimming. In The Physiology of Fishes (Evans, D. H., ed.), pp. 3–24.
New York: CRC Marine Science Series.
Webb, P. W. (2002). Control of posture, depth, and swimming trajectories of fishes.
Integrative and Comparative Biology 42, 94–101.
Webb, P. W. & Blake, R. W. (1985). Swimming. In Functional Vertebrate Morphology
(Hildebrand, M., Bramble, D. M., Liem, K. F. & Wake, D. B., eds), pp. 111–128.
Harvard: Belknap Press.
Webb, P. W. & de Buffrenil, V. (1990). Locomotion in the biology of large aquatic
vertebrates. Transactions of the American Fisheries Society 119, 629–641.
Webb, P. W. & Gerster, C. L. (2000). Swimming behaviour: predictions from hydromechanical
principles. In Biomechanics in Animal Behaviour (Domenici, P. D. & Blake, R. W.,
eds), pp. 59–77. Oxford: Bios Scientific Publishers Ltd.
Webb, P. W. & Keyes, R. S. (1981). Division of labour between median fins in swimming
dolphin (Pisces: Coryphaeidae). Copeia 1981, 901–904.
Webb, P. W. & Skadsen, J. M. (1980). Strike tactics of Esox. Canadian Journal of Zoology
58, 1462–1469.
Weihs, D. (1973a). The mechanism of rapid starting of slender fish. Biorheology 10, 343–350.
Weihs, D. (1973b). Optimal fish cruising speed. Nature 245, 48–50.
Weihs, D. (1973c). Mechanically efficient swimming techniques for fish with negative
buoyancy. Journal of Marine Research 31, 194–209.
Weihs, D. (1974). Some hydromechanical aspects of fish schooling. In Swimming and
Flying in Nature, Vol. 2 (Wu, T. Y., Brokaw, C. J. & Brennen, C., eds),
pp. 203–218. New York: Plenum Press.
Weihs, D. (1989). Design features and mechanics of axial locomotion in fish. American
Zoologist 29, 151–160.
Weihs, D. (2002). Stability versus maneuverability in aquatic locomotion. Integrative and
Comparative Biology 42, 127–134.
Weihs, D. & Webb, P. W. (1983). Optimization of locomotion. In Fish Biomechanics
(Webb, P. W. & Weihs, D., eds), pp. 339–371. New York: Praeger.
Wu, T. Y. (1971). Hydromechanics of swimming propulsion. Part 1. Swimming of a
two-dimensional flexible plate at variable forward speeds in an inviscid fluid.
Journal of Fluid Mechanics 46, 337–355.

# 2004 The Fisheries Society of the British Isles, Journal of Fish Biology 2004, 65, 1193–1222

You might also like