Fish locomotion

Last updated
Fish, like these yellowfin tuna, use many different mechanisms to propel themselves through water Yellowfin tuna nurp.jpg
Fish, like these yellowfin tuna, use many different mechanisms to propel themselves through water

Fish locomotion is the various types of animal locomotion used by fish, principally by swimming. This is achieved in different groups of fish by a variety of mechanisms of propulsion, most often by wave-like lateral flexions of the fish's body and tail in the water, and in various specialised fish by motions of the fins. The major forms of locomotion in fish are:

Contents

More specialized fish include movement by pectoral fins with a mainly stiff body, opposed sculling with dorsal and anal fins, as in the sunfish; and movement by propagating a wave along the long fins with a motionless body, as in the knifefish or featherbacks.

In addition, some fish can variously "walk" (i.e., crawl over land using the pectoral and pelvic fins), burrow in mud, leap out of the water and even glide temporarily through the air.

Swimming

Mechanism

Fins used for locomotion: (1) pectoral fins (paired), (2) pelvic fins (paired), (3) dorsal fin, (4) adipose fin, (5) anal fin, (6) caudal (tail) fin Lampanyctodes hectoris (fins).png
Fins used for locomotion: (1) pectoral fins (paired), (2) pelvic fins (paired), (3) dorsal fin, (4) adipose fin, (5) anal fin, (6) caudal (tail) fin

Fish swim by exerting force against the surrounding water. There are exceptions, but this is normally achieved by the fish contracting muscles on either side of its body in order to generate waves of flexion that travel the length of the body from nose to tail, generally getting larger as they go along. The vector forces exerted on the water by such motion cancel out laterally, but generate a net force backwards which in turn pushes the fish forward through the water. Most fishes generate thrust using lateral movements of their body and caudal fin, but many other species move mainly using their median and paired fins. The latter group swim slowly, but can turn rapidly, as is needed when living in coral reefs for example. But they can not swim as fast as fish using their bodies and caudal fins. [1] [2]

Skeletal anatomy of Tilapia Skeletal anatomy of tilapia.png
Skeletal anatomy of Tilapia

Consider the tilapia shown in the diagram. Like most fish, the tilapia has a streamlined body shape reducing water resistance to movement and enabling the tilapia to cut easily through water. Its head is inflexible, which helps it maintain forward thrust. [3] Its scales overlap and point backwards, allowing water to pass over the fish without unnecessary obstruction. Water friction is further reduced by mucus which tilapia secrete over their body. [3]

Like a plane or submarine, a fish has six degrees of freedom. 6DOF en.jpg
Like a plane or submarine, a fish has six degrees of freedom.

The backbone is flexible, allowing muscles to contract and relax rhythmically and bring about undulating movement. [3] A swim bladder provides buoyancy which helps the fish adjust its vertical position in the water column. A lateral line system allows it to detect vibrations and pressure changes in water, helping the fish to respond appropriately to external events. [3]

Well developed fins are used for maintaining balance, braking and changing direction. The pectoral fins act as pivots around which the fish can turn rapidly and steer itself. The paired pectoral and pelvic fins control pitching, while the unpaired dorsal and anal fins reduce yawing and rolling. The caudal fin provides raw power for propelling the fish forward. [3]

Body/caudal fin propulsion

There are five groups that differ in the fraction of their body that is displaced laterally: [1]

Anguilliform

Eels propagate a more or less constant-sized flexion wave along their slender bodies. FMIB 35739 Anguilla vulgaris -- Anguilla.jpeg
Eels propagate a more or less constant-sized flexion wave along their slender bodies.

In the anguilliform group, containing some long, slender fish such as eels, there is little increase in the amplitude of the flexion wave as it passes along the body. [1] [4]

Subcarangiform

The subcarangiform group has a more marked increase in wave amplitude along the body with the vast majority of the work being done by the rear half of the fish. In general, the fish body is stiffer, making for higher speed but reduced maneuverability. Trout use sub-carangiform locomotion. [1]

Carangiform

The carangiform group, named for the Carangidae, are stiffer and faster-moving than the previous groups. The vast majority of movement is concentrated in the very rear of the body and tail. Carangiform swimmers generally have rapidly oscillating tails. [1]

Thunniform

Tunas such as the bluefin swim fast with their large crescent-shaped tails. Bluefin-big.jpg
Tunas such as the bluefin swim fast with their large crescent-shaped tails.

The thunniform group contains high-speed long-distance swimmers, and is characteristic of tunas [5] and is also found in several lamnid sharks. [6] Here, virtually all the sideways movement is in the tail and the region connecting the main body to the tail (the peduncle). The tail itself tends to be large and crescent shaped. [1]

Ostraciiform

The ostraciiform group have no appreciable body wave when they employ caudal locomotion. Only the tail fin itself oscillates (often very rapidly) to create thrust. This group includes Ostraciidae. [1]

Median/paired fin propulsion

Boxfish use median-paired fin swimming, as they are not well streamlined, and use primarily their pectoral fins to produce thrust. Lactoria cornuta aka longhorn cowfish in cph aquarium 2007.jpg
Boxfish use median-paired fin swimming, as they are not well streamlined, and use primarily their pectoral fins to produce thrust.

Not all fish fit comfortably in the above groups. Ocean sunfish, for example, have a completely different system, the tetraodontiform mode, and many small fish use their pectoral fins for swimming as well as for steering and dynamic lift. Fish in the order Gymnotiformes possess electric organs along the length of their bodies and swim by undulating an elongated anal fin while keeping the body still, presumably so as not to disturb the electric field that they generate.

Many fish swim using combined behavior of their two pectoral fins or both their anal and dorsal fins. Different types of Median paired fin propulsion can be achieved by preferentially using one fin pair over the other, and include rajiform, diodontiform, amiiform, gymnotiform and balistiform modes. [2]

Rajiform


Rajiform locomotion is characteristic of rays and skates, when thrust is produced by vertical undulations along large, well developed pectoral fins. [2]

Diodontiform

Porcupine fish (here, Diodon nicthemerus) swim by undulating their pectoral fins. Diodon nicthemerus.jpg
Porcupine fish (here, Diodon nicthemerus ) swim by undulating their pectoral fins.

Diodontiform locomotion propels the fish propagating undulations along large pectoral fins, as seen in the porcupinefish (Diodontidae). [2]

Amiiform

Amiiform locomotion consists of undulations of a long dorsal fin while the body axis is held straight and stable, as seen in the bowfin. [2]

Gymnotiform

Gymnotus maintains a straight back while swimming to avoid disturbing its electric sense. Gymnotus sp.jpg
Gymnotus maintains a straight back while swimming to avoid disturbing its electric sense.

Gymnotiform locomotion consists of undulations of a long anal fin, essentially upside down amiiform, seen in the South American knifefish Gymnotiformes . [2]

Balistiform


In balistiform locomotion, both anal and dorsal fins undulate. It is characteristic of the family Balistidae (triggerfishes). It may also be seen in the Zeidae. [2]

Oscillatory

Oscillation is viewed as pectoral-fin-based swimming and is best known as mobuliform locomotion. The motion can be described as the production of less than half a wave on the fin, similar to a bird wing flapping. Pelagic stingrays, such as the manta, cownose, eagle and bat rays use oscillatory locomotion. [7]

Tetraodontiform

In tetraodontiform locomotion, the dorsal and anal fins are flapped as a unit, either in phase or exactly opposing one another, as seen in the Tetraodontiformes (boxfishes and pufferfishes). The ocean sunfish displays an extreme example of this mode. [2]

Labriform

In labriform locomotion, seen in the wrasses (Labriformes), oscillatory movements of pectoral fins are either drag based or lift based. Propulsion is generated either as a reaction to drag produced by dragging the fins through the water in a rowing motion, or via lift mechanisms. [2] [8]

Dynamic lift

Sharks are denser than water and must swim continually to maintain depth, using dynamic lift from their pectoral fins. Tiburon.jpg
Sharks are denser than water and must swim continually to maintain depth, using dynamic lift from their pectoral fins.

Bone and muscle tissues of fish are denser than water. To maintain depth, bony fish increase buoyancy by means of a gas bladder. Alternatively, some fish store oils or lipids for this same purpose. Fish without these features use dynamic lift instead. It is done using their pectoral fins in a manner similar to the use of wings by airplanes and birds. As these fish swim, their pectoral fins are positioned to create lift which allows the fish to maintain a certain depth. The two major drawbacks of this method are that these fish must stay moving to stay afloat and that they are incapable of swimming backwards or hovering. [9] [10]

Hydrodynamics

Similarly to the aerodynamics of flight, powered swimming requires animals to overcome drag by producing thrust. Unlike flying, however, swimming animals often do not need to supply much vertical force because the effect of buoyancy can counter the downward pull of gravity, allowing these animals to float without much effort. While there is great diversity in fish locomotion, swimming behavior can be classified into two distinct "modes" based on the body structures involved in thrust production, Median-Paired Fin (MPF) and Body-Caudal Fin (BCF). Within each of these classifications, there are numerous specifications along a spectrum of behaviours from purely undulatory to entirely oscillatory. In undulatory swimming modes, thrust is produced by wave-like movements of the propulsive structure (usually a fin or the whole body). Oscillatory modes, on the other hand, are characterized by thrust produced by swiveling of the propulsive structure on an attachment point without any wave-like motion. [2]

Body-caudal fin

Sardines use body-caudal fin propulsion to swim, holding their pectoral, dorsal, and anal fins flat against the body, creating a more streamlined body to reduce drag.

Most fish swim by generating undulatory waves that propagate down the body through the caudal fin. This form of undulatory locomotion is termed body-caudal fin (BCF) swimming on the basis of the body structures used; it includes anguilliform, sub-carangiform, carangiform, and thunniform locomotory modes, as well as the oscillatory ostraciiform mode. [2] [11]

Adaptation

Similar to adaptation in avian flight, swimming behaviors in fish can be thought of as a balance of stability and maneuverability. [12] Because body-caudal fin swimming relies on more caudal body structures that can direct powerful thrust only rearwards, this form of locomotion is particularly effective for accelerating quickly and cruising continuously. [2] [11] body-caudal fin swimming is, therefore, inherently stable and is often seen in fish with large migration patterns that must maximize efficiency over long periods. Propulsive forces in median-paired fin swimming, on the other hand, are generated by multiple fins located on either side of the body that can be coordinated to execute elaborate turns. As a result, median-paired fin swimming is well adapted for high maneuverability and is often seen in smaller fish that require elaborate escape patterns. [12]

The habitats occupied by fishes are often related to their swimming capabilities. On coral reefs, the faster-swimming fish species typically live in wave-swept habitats subject to fast water flow speeds, while the slower fishes live in sheltered habitats with low levels of water movement. [13]

Fish do not rely exclusively on one locomotor mode, but are rather locomotor generalists, [2] choosing among and combining behaviors from many available behavioral techniques. Predominantly body-caudal fin swimmers often incorporate movement of their pectoral, anal, and dorsal fins as an additional stabilizing mechanism at slower speeds, [14] but hold them close to their body at high speeds to improve streamlining and reducing drag. [2] Zebrafish have even been observed to alter their locomotor behavior in response to changing hydrodynamic influences throughout growth and maturation. [15]

Flight

The transition of predominantly swimming locomotion directly to flight has evolved in a single family of marine fish, the Exocoetidae. Flying fish are not true fliers in the sense that they do not execute powered flight. Instead, these species glide directly over the surface of the ocean water without ever flapping their "wings." Flying fish have evolved abnormally large pectoral fins that act as airfoils and provide lift when the fish launches itself out of the water. Additional forward thrust and steering forces are created by dipping the hypocaudal (i.e. bottom) lobe of their caudal fin into the water and vibrating it very quickly, in contrast to diving birds in which these forces are produced by the same locomotor module used for propulsion. Of the 64 extant species of flying fish, only two distinct body plans exist, each of which optimizes two different behaviors. [16] [17]

Flying fish gain sufficient lift to glide above the water thanks to their enlarged pectoral fins. Schwalbenfisch.jpg
Flying fish gain sufficient lift to glide above the water thanks to their enlarged pectoral fins.

Tradeoffs

While most fish have caudal fins with evenly sized lobes (i.e. homocaudal), flying fish have an enlarged ventral lobe (i.e. hypocaudal) which facilitates dipping only a portion of the tail back onto the water for additional thrust production and steering. [17]

Because flying fish are primarily aquatic animals, their body density must be close to that of water for buoyancy stability. This primary requirement for swimming, however, means that flying fish are heavier (have a larger mass) than other habitual fliers, resulting in higher wing loading and lift to drag ratios for flying fish compared to a comparably sized bird. [16] Differences in wing area, wing span, wing loading, and aspect ratio have been used to classify flying fish into two distinct classifications based on these different aerodynamic designs. [16]

Biplane body plan

In the biplane or Cypselurus body plan, both the pectoral and pelvic fins are enlarged to provide lift during flight. [16] These fish also tend to have "flatter" bodies which increase the total lift-producing area, thus allowing them to "hang" in the air better than more streamlined shapes. [17] As a result of this high lift production, these fish are excellent gliders and are well adapted for maximizing flight distance and duration.

Comparatively, Cypselurus flying fish have lower wing loading and smaller aspect ratios (i.e. broader wings) than their Exocoetus monoplane counterparts, which contributes to their ability to fly for longer distances than fish with this alternative body plan. Flying fish with the biplane design take advantage of their high lift production abilities when launching from the water by utilizing a "taxiing glide" in which the hypocaudal lobe remains in the water to generate thrust even after the trunk clears the water's surface and the wings are opened with a small angle of attack for lift generation. [16]

In the monoplane body plan of Exocoetus, only the pectoral fins are abnormally large, while the pelvic fins are small. Sailfin flyingfish.jpg
In the monoplane body plan of Exocoetus , only the pectoral fins are abnormally large, while the pelvic fins are small.

Monoplane body plan

In the Exocoetus or monoplane body plan, only the pectoral fins are enlarged to provide lift. Fish with this body plan tend to have a more streamlined body, higher aspect ratios (long, narrow wings), and higher wing loading than fish with the biplane body plan, making these fish well adapted for higher flying speeds. Flying fish with a monoplane body plan demonstrate different launching behaviors from their biplane counterparts. Instead of extending their duration of thrust production, monoplane fish launch from the water at high speeds at a large angle of attack (sometimes up to 45 degrees). [16] In this way, monoplane fish are taking advantage of their adaptation for high flight speed, while fish with biplane designs exploit their lift production abilities during takeoff.

Walking

Alticus arnoldorum hopping
Alticus arnoldorum climbing up a vertical piece of Plexiglas

A "walking fish" is a fish that is able to travel over land for extended periods of time. Some other cases of nonstandard fish locomotion include fish "walking" along the sea floor, such as the handfish or frogfish.

Most commonly, walking fish are amphibious fish. Able to spend longer times out of water, these fish may use a number of means of locomotion, including springing, snake-like lateral undulation, and tripod-like walking. The mudskippers are probably the best land-adapted of contemporary fish and are able to spend days moving about out of water and can even climb mangroves, although to only modest heights. [18] The Climbing gourami is often specifically referred to as a "walking fish", although it does not actually "walk", but rather moves in a jerky way by supporting itself on the extended edges of its gill plates and pushing itself by its fins and tail. Some reports indicate that it can also climb trees. [19]

There are a number of fish that are less adept at actual walking, such as the walking catfish. Despite being known for "walking on land", this fish usually wriggles and may use its pectoral fins to aid in its movement. Walking Catfish have a respiratory system that allows them to live out of water for several days. Some are invasive species. A notorious case in the United States is the Northern snakehead. [20] Polypterids have rudimentary lungs and can also move about on land, though rather clumsily. The Mangrove rivulus can survive for months out of water and can move to places like hollow logs. [21] [22] [23] [24]

Ogcocephalus parvus OgcocephalusParvus.jpg
Ogcocephalus parvus

There are some species of fish that can "walk" along the sea floor but not on land; one such animal is the flying gurnard (it does not actually fly, and should not be confused with flying fish). The batfishes of the family Ogcocephalidae (not to be confused with batfish of Ephippidae) are also capable of walking along the sea floor. Bathypterois grallator , also known as a "tripodfish", stands on its three fins on the bottom of the ocean and hunts for food. [25] The African lungfish (P. annectens) can use its fins to "walk" along the bottom of its tank in a manner similar to the way amphibians and land vertebrates use their limbs on land. [26] [27] [28]

Burrowing

Many fishes, particularly eel-shaped fishes such as true eels, moray eels, and spiny eels, are capable of burrowing through sand or mud. [29] Ophichthids, the snake eels, are capable of burrowing either forwards or backwards. [30]

In larvae

Swimming

Salmon larva emerging from its egg Salmonlarvakils.jpg
Salmon larva emerging from its egg

Fish larvae, like many adult fishes, swim by undulating their body. The swimming speed varies proportionally with the size of the animals, in that smaller animals tend to swim at lower speeds than larger animals. The swimming mechanism is controlled by the flow regime of the larvae. Reynolds number (Re) is defined as the ratio of inertial force to viscous force. Smaller organisms are affected more by viscous forces, like friction, and swim at a smaller Reynolds number. Larger organisms use a larger proportion of inertial forces, like pressure, to swim, at a higher Reynolds number. [31]

The larvae of ray finned fishes, the Actinopterygii, swim at a quite large range of Reynolds number (Re ≈10 to 900). This puts them in an intermediate flow regime where both inertial and viscous forces play an important role. As the size of the larvae increases, the use of pressure forces to swim at higher Reynolds number increases.

Undulatory swimmers generally shed at least two types of wake: Carangiform swimmers shed connected vortex loops and Anguilliform swimmers shed individual vortex rings. These vortex rings depend upon the shape and arrangement of the trailing edge from which the vortices are shed. These patterns depend upon the swimming speed, ratio of swimming speed to body wave speed and the shape of body wave. [31]

A spontaneous bout of swimming has three phases. The first phase is the start or acceleration phase: In this phase the larva tends to rotate its body to make a 'C' shape which is termed the preparatory stroke. It then pushes in the opposite direction to straighten its body, which is called a propulsive stroke, or a power stroke, which powers the larva to move forward. The second phase is cyclic swimming. In this phase, the larva swims with an approximately constant speed. The last phase is deceleration. In this phase, the swimming speed of the larva gradually slows down to a complete stop. In the preparatory stroke, due to the bending of the body, the larva creates 4 vortices around its body, and 2 of those are shed in the propulsive stroke. [31] Similar phenomena can be seen in the deceleration phase. However, in the vortices of the deceleration phase, a large area of elevated vorticity can be seen compared to the starting phase.

The swimming abilities of larval fishes are important for survival. This is particularly true for the larval fishes with higher metabolic rate and smaller size which makes them more susceptible to predators. The swimming ability of a reef fish larva helps it to settle at a suitable reef and for locating its home as it is often isolated from its home reef in search of food. Hence the swimming speed of reef fish larvae are quite high (≈12 cm/s - 100 cm/s) compared to other larvae. [32] [33] The swimming speeds of larvae from the same families at the two locations are relatively similar. [32] However, the variation among individuals is quite large. At the species level, length is significantly related to swimming ability. However, at the family level, only 16% of variation in swimming ability can be explained by length. [32] There is also a negative correlation between the fineness ratio (length of body to maximum width) and the swimming ability of reef fish larvae. This suggests a minimization of overall drag and maximization of volume. Reef fish larvae differ significantly in their critical swimming speed abilities among taxa which leads to high variability in sustainable swimming speed. [34] This again leads to sustainable variability in their ability to alter dispersal patterns, overall dispersal distances and control their temporal and spatial patterns of settlement. [35]

Hydrodynamics

Small undulatory swimmers such as fish larvae experience both inertial and viscous forces, the relative importance of which is indicated by Reynolds number (Re). Reynolds number is proportional to body size and swimming speed. The swimming performance of a larva increases between 2–5 days post fertilization. Compared with adults, larval fish experience relatively high viscous force. To enhance thrust to an equal level with the adults, it increases its tail beat frequency and thus amplitude. In zebrafish, tail beat frequency increases over larval age to 95 Hz in 3 days post fertilization from 80 Hz in 2 days post fertilization. This higher frequency leads to higher swimming speed, thus reducing predation and increasing prey catching ability when they start feeding at around 5 days post fertilization. The vortex shedding mechanics changes with the flow regime in an inverse non-linear way. Strouhal number is a design parameter for the vortex shedding mechanism. It can be defined as a ratio of the product of tail beat frequency with amplitude with the mean swimming speed. [36] Reynolds number (Re) is the main deciding criteria of a flow regime. It has been observed over different type of larval experiments that, slow larvae swims at higher Strouhal number but lower Reynolds number. However, the faster larvae swims distinctively at opposite conditions, that is, at lower Strouhal number but higher Reynolds number. Strouhal number is constant over similar speed ranged adult fishes. Strouhal number does not only depend on the small size of the swimmers, but also dependent to the flow regime. As in fishes which swim in viscous or high-friction flow regime, would create high body drag which will lead to higher Strouhal number. Whereas, in high viscous regime, the adults swim at lower stride length which leads to lower tail beat frequency and lower amplitude. This leads to higher thrust for same displacement or higher propulsive force, which unanimously reduces the Reynolds number. [37]

Larval fishes start feeding at 5–7 days post fertilization. And they experience extreme mortality rate (≈99%) in the few days after feeding starts. The reason for this 'Critical Period' (Hjort-1914) is mainly hydrodynamic constraints. Larval fish fail to eat even if there are enough prey encounters. One of the primary determinants of feeding success is the size of larval body. The smaller larvae function in a lower Reynolds number (Re) regime. As the age increases, the size of the larvae increases, which leads to higher swimming speed and increased Reynolds number. It has been observed through many experiments that the Reynolds number of successful strikes (Re~200) is much higher than the Reynolds number of failed strikes (Re~20). [38] [39] Numerical analysis of suction feeding at a low Reynolds number concluded that around 40% energy invested in mouth opening is lost to frictional forces rather than contributing to accelerating the fluid towards mouth. [40] Ontogenetic improvement in the sensory system, coordination and experiences are non-significant relationship while determining feeding success of larvae [39] A successful strike positively depends upon the peak flow speed or the speed of larvae at the time of strike. The peak flow speed is also dependent on the gape speed or the speed of opening the buccal cavity to capture food. As the larva ages, its body size increase and its gape speed also increase, which cumulatively increase the successful strike outcomes. [39]

The ability of a larval prey to survive an encounter with predator totally depends on its ability to sense and evade the strike. Adult fishes exhibit rapid suction feeding strikes as compared to larval fishes. Sensitivity of larval fish to velocity and flow fields provides the larvae a critical defense against predation. Though many prey use their visual system to detect and evade predators when there is light, it is hard for the prey to detect predators at night, which leads to a delayed response to the attack. There is a mechano-sensory system in fishes to identify the different flow generated by different motion surrounding the water and between the bodies called as lateral line system. [41] After detecting a predator, a larva evades its strike by 'fast start' or 'C' response. A swimming fish disturbs a volume of water ahead of its body with a flow velocity that increases with the proximity to the body. This particular phenomenon is sometimes called a bow wave. [42] The timing of the 'C' start response affects escape probability inversely. Escape probability increases with the distance from the predator at the time of strike. In general, prey successfully evade a predator strike from an intermediate distance (3–6 mm) from the predator. [41]

Behavior

Objective quantification is complicated in higher vertebrates by the complex and diverse locomotor repertoire and neural system. However, the relative simplicity of a juvenile brain and simple nervous system of fishes with fundamental neuronal pathways allows zebrafish larvae to be an apt model to study the interconnection between locomotor repertoire and neuronal system of a vertebrate. Behavior represents the unique interface between intrinsic and extrinsic forces that determine an organism's health and survival. [43] Larval zebrafish perform many locomotor behavior such as escape response, prey tracking, optomotor response etc. These behaviors can be categorized with respect to body position as ‘C’-starts, ‘J’-turns, slow scoots, routine turns etc. Fish larvae respond to abrupt changes in illumination with distinct locomotor behavior. The larvae show high locomotor activity during periods of bright light compared to dark. This behavior can direct towards the idea of searching food in light whereas the larvae do not feed in dark. [44] Also light exposure directly manipulates the locomotor activities of larvae throughout circadian period of light and dark with higher locomotor activity in light condition than in dark condition which is very similar as seen in mammals. Following the onset of darkness, larvae shows hyperactive scoot motion prior to a gradual drop off. This behavior could possibly be linked to find a shelter before nightfall. Also larvae can treat this sudden nightfall as under debris and the hyperactivity can be explained as the larvae navigation back to illuminated areas. [44] Prolonged dark period can reduce the light-dark responsiveness of larvae. Following light extinction, larvae execute large angle turns towards the vanished light source, which explains the navigational response of a larva. [44] Acute ethanol exposure reduce visual sensitivity of larvae causing a latency to respond in light and dark period change. [43]

See also

Related Research Articles

<span class="mw-page-title-main">Fin</span> Thin component or appendage attached to a larger body or structure

A fin is a thin component or appendage attached to a larger body or structure. Fins typically function as foils that produce lift or thrust, or provide the ability to steer or stabilize motion while traveling in water, air, or other fluids. Fins are also used to increase surface areas for heat transfer purposes, or simply as ornamentation.

<span class="mw-page-title-main">Gymnotiformes</span> Order of bony fishes

The Gymnotiformes are an order of teleost bony fishes commonly known as Neotropical knifefish or South American knifefish. They have long bodies and swim using undulations of their elongated anal fin. Found almost exclusively in fresh water, these mostly nocturnal fish are capable of producing electric fields to detect prey, for navigation, communication, and, in the case of the electric eel, attack and defense. A few species are familiar to the aquarium trade, such as the black ghost knifefish, the glass knifefish, and the banded knifefish.

<span class="mw-page-title-main">Teleost</span> Infraclass of fishes

Teleostei, members of which are known as teleosts, is, by far, the largest infraclass in the class Actinopterygii, the ray-finned fishes, and contains 96% of all extant species of fish. Teleosts are arranged into about 40 orders and 448 families. Over 26,000 species have been described. Teleosts range from giant oarfish measuring 7.6 m (25 ft) or more, and ocean sunfish weighing over 2 t, to the minute male anglerfish Photocorynus spiniceps, just 6.2 mm (0.24 in) long. Including not only torpedo-shaped fish built for speed, teleosts can be flattened vertically or horizontally, be elongated cylinders or take specialised shapes as in anglerfish and seahorses.

<span class="mw-page-title-main">Animal locomotion</span> Self-propulsion by an animal

Animal locomotion, in ethology, is any of a variety of methods that animals use to move from one place to another. Some modes of locomotion are (initially) self-propelled, e.g., running, swimming, jumping, flying, hopping, soaring and gliding. There are also many animal species that depend on their environment for transportation, a type of mobility called passive locomotion, e.g., sailing, kiting (spiders), rolling or riding other animals (phoresis).

<span class="mw-page-title-main">Jet propulsion</span> Thrust produced by ejecting a jet of fluid

Jet propulsion is the propulsion of an object in one direction, produced by ejecting a jet of fluid in the opposite direction. By Newton's third law, the moving body is propelled in the opposite direction to the jet. Reaction engines operating on the principle of jet propulsion include the jet engine used for aircraft propulsion, the pump-jet used for marine propulsion, and the rocket engine and plasma thruster used for spacecraft propulsion. Underwater jet propulsion is also used by several marine animals, including cephalopods and salps, with the flying squid even displaying the only known instance of jet-powered aerial flight in the animal kingdom.

<span class="mw-page-title-main">Rajiformes</span> Order of fishes in the superorder Batoidea

Rajiformes is one of the four orders in the superorder Batoidea, flattened cartilaginous fishes related to sharks. Rajiforms are distinguished by the presence of greatly enlarged pectoral fins, which reach as far forward as the sides of the head, with a generally flattened body. The undulatory pectoral fin motion diagnostic to this taxon is known as rajiform locomotion. The eyes and spiracles are located on the upper surface of the head and the gill slits are on the underside of the body. Most species give birth to live young, although some lay eggs enclosed in a horny capsule.

This glossary of ichthyology is a list of definitions of terms and concepts used in ichthyology, the study of fishes.

<span class="mw-page-title-main">Shark anatomy</span>

Shark anatomy differs from that of bony fish in a variety of ways. Variation observed within shark anatomy is a potential result of speciation and habitat variation.

<span class="mw-page-title-main">Epaulette shark</span> Species of longtailed carpet shark

The epaulette shark is a species of longtailed carpet shark of the family Hemiscylliidae, found in shallow, tropical waters off Australia and New Guinea. The common name of this shark comes from the very large, white-margined black spot behind each pectoral fin, which are reminiscent of military epaulettes. A small species usually under 1 m (3.3 ft) long, the epaulette shark has a slender body with a short head and broad, paddle-shaped paired fins. The caudal peduncle comprises over half the shark's length. Adults are light brown above, with scattered darker spots and indistinct saddles.

<i>Chromis atripectoralis</i> Species of fish

The black-axil chromis, also known as the blackfin chromis or blue-green puller, is a damselfish from the tropical Indo-Pacific. This fish can reach almost 12 centimetres (4.7 in) in length. It inhabits lagoons and reefs after settlement in the late-larval stage, and often occurs in large numbers, feeding above Acropora corals. This fish mostly feeds upon copepods, amphipods, and zoea.

Myomeres are blocks of skeletal muscle tissue arranged in sequence, commonly found in aquatic chordates. Myomeres are separated from adjacent myomeres by connective fascia (myosepta) and most easily seen in larval fishes or in the olm. Myomere counts are sometimes used for identifying specimens, since their number corresponds to the number of vertebrae in the adults. Location varies, with some species containing these only near the tails, while some have them located near the scapular or pelvic girdles. Depending on the species, myomeres could be arranged in an epaxial or hypaxial manner. Hypaxial refers to ventral muscles and related structures while epaxial refers to more dorsal muscles. The horizontal septum divides these two regions in vertebrates from cyclostomes to gnathostomes. In terrestrial chordates, the myomeres become fused as well as indistinct, due to the disappearance of myosepta.

<i>Histiophryne</i> Genus of fishes

Histiophryne is a genus of frogfishes found in waters ranging from Taiwan to South Australia. There are currently five known species. These fishes are easily distinguished from other anglerfishes as having a reduced luring appendage, a highly evolved form of the first dorsal fin spine.

<span class="mw-page-title-main">Aquatic locomotion</span>

Aquatic locomotion or swimming is biologically propelled motion through a liquid medium. The simplest propulsive systems are composed of cilia and flagella. Swimming has evolved a number of times in a range of organisms including arthropods, fish, molluscs, amphibians, reptiles, birds, and mammals.

<span class="mw-page-title-main">Fin and flipper locomotion</span>

Fin and flipper locomotion occurs mostly in aquatic locomotion, and rarely in terrestrial locomotion. From the three common states of matter — gas, liquid and solid, these appendages are adapted for liquids, mostly fresh or saltwater and used in locomotion, steering and balancing of the body. Locomotion is important in order to escape predators, acquire food, find mates and bury for shelter, nest or food. Aquatic locomotion consists of swimming, whereas terrestrial locomotion encompasses walking, 'crutching', jumping, digging as well as covering. Some animals such as sea turtles and mudskippers use these two environments for different purposes, for example using the land for nesting, and the sea to hunt for food.

<span class="mw-page-title-main">Undulatory locomotion</span>

Undulatory locomotion is the type of motion characterized by wave-like movement patterns that act to propel an animal forward. Examples of this type of gait include crawling in snakes, or swimming in the lamprey. Although this is typically the type of gait utilized by limbless animals, some creatures with limbs, such as the salamander, forgo use of their legs in certain environments and exhibit undulatory locomotion. In robotics this movement strategy is studied in order to create novel robotic devices capable of traversing a variety of environments.

<span class="mw-page-title-main">Gray's paradox</span>

Gray's Paradox is a paradox posed in 1936 by British zoologist Sir James Gray. The paradox was to figure out how dolphins can obtain such high speeds and accelerations with what appears to be a small muscle mass. Gray made an estimate of the power a dolphin could exert based on its physiology, and concluded the power was insufficient to overcome the drag forces in water. He hypothesized that Dolphin's skin must have special anti-drag properties.

<span class="mw-page-title-main">Tradeoffs for locomotion in air and water</span> Comparison of swimming and flying, evolution and biophysics

Certain species of fish and birds are able to locomote in both air and water, two fluid media with very different properties. A fluid is a particular phase of matter that deforms under shear stresses and includes any type of liquid or gas. Because fluids are easily deformable and move in response to applied forces, efficiently locomoting in a fluid medium presents unique challenges. Specific morphological characteristics are therefore required in animal species that primarily depend on fluidic locomotion. Because the properties of air and water are so different, swimming and flying have very disparate morphological requirements. As a result, despite the large diversity of animals that are capable of flight or swimming, only a limited number of these species have mastered the ability to both fly and swim. These species demonstrate distinct morphological and behavioral tradeoffs associated with transitioning from air to water and water to air.

<span class="mw-page-title-main">Fish fin</span> Bony skin-covered spines or rays protruding from the body of a fish

Fins are distinctive anatomical features composed of bony spines or rays protruding from the body of Actinopterygii, Dipnomorpha, Actinistia and Chondrichthyes fishes. They are covered with skin and joined together either in a webbed fashion, as seen in most bony fish, or similar to a flipper, as seen in sharks. Apart from the tail or caudal fin, fish fins have no direct connection with the spine and are supported only by muscles. Their principal function is to help the fish swim.

<span class="mw-page-title-main">Robot fish</span> Robot designed to move like a living fish

A robot fish is a type of bionic robot that has the shape and locomotion of a living fish. Most robot fish are designed to emulate living fish which use body-caudal fin (BCF) propulsion, and can be divided into three categories: single joint (SJ), multi-joint (MJ) and smart material-based "soft-body" design.

Batoids are a superorder of cartilaginous fish consisting of skates, rays and other fish all characterized by dorsoventrally flattened bodies and large pectoral fins fused to the head. This distinctive morphology has resulted in several unique forms of locomotion. Most Batoids exhibit median paired fin swimming, utilizing their enlarged pectoral fins. Batoids that exhibit median paired fin swimming fall somewhere along a spectrum of swimming modes from mobuliform to rajiform based on the number of waves present on their fin at once. Of the four orders of Batoidae this holds truest for the Myliobatiformes (rays) and the Rajiformes (skates). The two other orders: Rhinopristiformes and Torpediniformes exhibit a greater degree of body caudal fin swimming.

References

  1. 1 2 3 4 5 6 7 Breder, CM (1926). "The locomotion of fishes". Zoologica. 4: 159–297.
  2. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Sfakiotakis, M.; Lane, D. M.; Davies, J. B. C. (1999). "Review of Fish Swimming Modes for Aquatic Locomotion" (PDF). IEEE Journal of Oceanic Engineering. 24 (2): 237–252. Bibcode:1999IJOE...24..237S. doi:10.1109/48.757275. S2CID   17226211. Archived from the original (PDF) on 2013-12-24.
  3. 1 2 3 4 5 6 Locomotion in Finned Fish, Global e-Schools and Communities Initiative (GeSCI) United Nations. Retrieved 7 Sep 2021. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  4. Long Jr, J. H., Shepherd, W., & Root, R. G. (1997). Manueuverability and reversible propulsion: How eel-like fish swim forward and backward using travelling body waves". In: Proc. Special Session on Bio-Engineering Research Related to Autonomous Underwater Vehicles, 10th Int. Symp. Unmanned Untethered Submersible Technology (pp. 118-134).
  5. Hawkins, JD; Sepulveda, CA; Graham, JB; Dickson, KA (2003). "Swimming performance studies on the eastern Pacific bonito Sarda chiliensis, a close relative of the tunas (family Scombridae) II. Kinematics". The Journal of Experimental Biology. 206 (16): 2749–2758. doi: 10.1242/jeb.00496 . PMID   12847120.
  6. Klimley, A. Peter (2013). The Biology of Sharks, Skates, and Rays. University of Chicago Press. ISBN   978-0-226-44249-5.
  7. Lindsey, C.C. (1978). "Locomotion". In Hoar W.S.; Randall, D.J. (eds.). Fish Physiology. Vol. 7. Academic Press. San Francisco. pp. 1–100.
  8. Fulton, CJ; Johansen, JL; Steffensen, JF (2013). "Energetic extremes in aquatic locomotion by coral reef fishes". PLOS ONE. 8 (1): e54033. Bibcode:2013PLoSO...854033F. doi: 10.1371/journal.pone.0054033 . PMC   3541231 . PMID   23326566.
  9. Bennetta, William J. (1996). "Deep Breathing". Archived from the original on 2007-08-14. Retrieved 2007-08-28.
  10. "Do sharks sleep". Flmnh.ufl.edu. 2017-05-02. Archived from the original on 2010-09-18.
  11. 1 2 Blake, R. W. (2004). "Review Paper: Fish functional design and swimming performance". Journal of Fish Biology. 65 (5): 1193–1222. doi:10.1111/j.0022-1112.2004.00568.x.
  12. 1 2 Weihs, Daniel (2002). "Stability versus Maneuverability in Aquatic Locomotion". Integrated and Computational Biology. 42 (1): 127–134. doi: 10.1093/icb/42.1.127 . PMID   21708701.
  13. Fulton, C. J.; Bellwood, D. R.; Wainwright, P. C. (2005). "Wave energy and swimming performance shape coral reef fish assemblages". Proceedings of the Royal Society B. 272 (1565): 827–832. doi:10.1098/rspb.2004.3029. PMC   1599856 . PMID   15888415.
  14. Heatwole, S. J.; Fulton, C. J. (2013). "Behavioural flexibility in coral reef fishes responding to a rapidly changing environment". Marine Biology. 160 (3): 677–689. doi:10.1007/s00227-012-2123-2. S2CID   85119253.
  15. McHenry, Matthew J.; Lauder, George V. (2006). "Ontogeny of Form and Function: Locomotor Morphology and Drag in Zebrafish (Danio rerio)". Journal of Morphology. 267 (9): 1099–1109. doi:10.1002/jmor.10462. PMID   16752407. S2CID   33343483.
  16. 1 2 3 4 5 6 Fish, F.E. (1990) Wing design and scaling of flying fish with regard to flight performance. "J. Zool. Lond." 221, 391-403.
  17. 1 2 3 Fish, Frank. (1991) On a Fin and a Prayer. "Scholars." 3(1), 4-7.
  18. "Cairns Museum Tour - Cairns-Kuranda Railway". Archived from the original on 2015-01-08. Retrieved 2015-01-08.
  19. "Climbing Fish". Archived from the original on 2009-08-29. Retrieved 2015-02-26.
  20. "Maryland Suffers Setback in War on Invasive Walking Fish", National Geographic News July 12, 2002
  21. Shells, trees and bottoms: Strange places fish live
  22. "Tropical fish can live for months out of water". Reuters. 15 November 2007.
  23. Fish Lives in Logs, Breathing Air, for Months at a Time
  24. Fish Lives in Logs, Breathing Air, for Months at a Time
  25. Jones, AT; KJ Sulak (1990). "First Central Pacific Plate and Hawaiian Record of the Deep-sea Tripod Fish Bathypterois grallator (Pisces: Chlorophthalmidae)" (PDF). Pacific Science. 44 (3): 254–7.
  26. Fish uses fins to walk and bound
  27. Behavioral evidence for the evolution of walking and bounding before terrestriality in sarcopterygian fishes
  28. A Small Step for Lungfish, a Big Step for the Evolution of Walking
  29. Monks, Neale (2006). Brackish-Water Fishes. TFH. pp. 223–226. ISBN   978-0-7938-0564-8.
  30. Allen, Gerry (1999). Marine Fishes of Southeast Asia: A Field Guide for Anglers and Divers. Tuttle Publishing. p. 56. ISBN   978-1-4629-1707-5. many have a bony, sharp tail and are equally adept at burrowing forward or backward.
  31. 1 2 3 ‘Flow Patterns Of Larval Fish: Undulatory Swimming in the Intermediate Flow Regime’ by Ulrike K. Müller, Jos G. M. van den Boogaart and Johan L. van Leeuwen. Journal of Experimental Biology 2008 211: 196-205; doi: 10.1242/jeb.005629
  32. 1 2 3 "Critical Swimming Speeds of Late-Stage Coral Reef Fish Larvae: Variation within Species, Among Species and Between Locations" by Fisher, R., Leis, J.M., Clark, D.L.in Marine Biology (2005) 147: 1201. https://doi.org/10.1007/s00227-005-0001-x,
  33. "Development of Swimming Abilities in Reef Fish Larvae" by Rebecca Fisher, David R. Bellwood, Suresh D. Job in Marine Ecology-progress Series - MAR ECOL-PROGR SER. 202. 163-173. 10.3354/meps202163
  34. ‘Maximum Sustainable Swimming Speeds Of Late-Stage Larvae Of Nine Species Of Reef Fishes’ by Rebecca Fisher, Shaun K.Wilson in Journal of Experimental Marine Biology and Ecology, Volume 312, Issue 1, 2004, Pages 171-186, ISSN 0022-0981, https://doi.org/10.1016/j.jembe.2004.06.009
  35. 'Development of Swimming Abilities in Reef Fish Larvae' by Rebecca Fisher, David R. Bellwood, Suresh D. Job in Marine Ecology-progress Series - MAR ECOL-PROGR SER. 202. 163-173. 10.3354/meps202163
  36. van Leeuwen, Johan L.; Voesenek, Cees J.; Müller, Ulrike K. (2015). "How body torque and Strouhal number change with swimming speed and developmental stage in larval zebrafish". Journal of the Royal Society Interface. The Royal Society. 12 (110): 20150479. doi: 10.1098/rsif.2015.0479 . ISSN   1742-5689. PMC   4614456 . PMID   26269230.
  37. 'How body torque and Strouhal number change with swimming speed and developmental stage in larval zebrafish' by Johan L. van Leeuwen, Cees J. Voesenek and Ulrike K. Müller in J. R. Soc. Interface 2015 12 20150479; DOI: 10.1098/rsif.2015.0479. 6 September 2015
  38. China, Victor; Holzman, Roi (19 May 2014). "Hydrodynamic starvation in first-feeding larval fishes". Proceedings of the National Academy of Sciences. 111 (22): 8083–8088. Bibcode:2014PNAS..111.8083C. doi: 10.1073/pnas.1323205111 . ISSN   0027-8424. PMC   4050599 . PMID   24843180.
  39. 1 2 3 China, Victor; Levy, Liraz; Liberzon, Alex; Elmaliach, Tal; Holzman, Roi (26 April 2017). "Hydrodynamic regime determines the feeding success of larval fish through the modulation of strike kinematics". Proceedings of the Royal Society B: Biological Sciences. The Royal Society. 284 (1853): 20170235. doi: 10.1098/rspb.2017.0235 . ISSN   0962-8452. PMC   5413926 . PMID   28446697.
  40. Drost, M. R.; Muller, M.; Osse, J. W. M. (23 August 1988). "A quantitative hydrodynamical model of suction feeding in larval fishes: the role of frictional forces". Proceedings of the Royal Society of London. Series B. Biological Sciences. The Royal Society. 234 (1276): 263–281. Bibcode:1988RSPSB.234..263D. doi:10.1098/rspb.1988.0048. ISSN   0080-4649. S2CID   86188901.
  41. 1 2 Stewart, William J.; Cardenas, Gilberto S.; McHenry, Matthew J. (1 February 2013). "Zebrafish larvae evade predators by sensing water flow". Journal of Experimental Biology. The Company of Biologists. 216 (3): 388–398. doi: 10.1242/jeb.072751 . ISSN   1477-9145. PMID   23325859.
  42. Ferry-Graham, Lara A.; Wainwright, Peter C.; Lauder, George V. (2003). "Quantification of flow during suction feeding in bluegill sunfish". Zoology. Elsevier. 106 (2): 159–168. doi:10.1078/0944-2006-00110. ISSN   0944-2006. PMID   16351901.
  43. 1 2 ‘Locomotion In Larval Zebrafish: Influence of Time of Day, Lighting and Ethanol’ by R.C. MacPhail, J. Brooks, D.L. Hunter, B. Padnos a, T.D. Irons, S. Padilla in Neurotoxicology. 30. 52-8. 10.1016/j.neuro.2008.09.011.
  44. 1 2 3 ‘Modulation of Locomotor Activity in Larval Zebrafish During Light Adaptation’ by Harold A. Burgess and Michael Granato. In Journal of Experimental Biology 2007 210: 2526-2539; doi: 10.1242/jeb.003939

Further reading