locomotion

locomotion
/loh'keuh moh"sheuhn/, n.
the act or power of moving from place to place.
[1640-50; see LOCOMOTIVE, MOTION]

* * *

Any of various animal movements that result in progression from one place to another.

Locomotion is classified as either appendicular (accomplished by special appendages) or axial (achieved by changing the body shape). Aquatic protozoans move by ciliary or flagellar appendages or by pseudopods, footlike appendages. Other forms of aquatic locomotion include walking on legs (some arthropods), crawling (by contracting the body muscles, anchoring to the substrate, and extending), and swimming by either hydraulic propulsion (e.g., jellyfish) or undulation (fishes). Terrestrial arthropods and vertebrates move by means of jointed appendages, the legs. Snakes and other limbless vertebrates crawl by means of muscular thrusts against the substrate. Flight is achieved by the forward thrust of wings.

* * *

Introduction

      in ethology, any of a variety of movements among animals that results in progression from one place to another.

Principles
      To locomote, all animals require both propulsive and control mechanisms. The diverse propulsive mechanisms of animals involve a contractile structure—muscle in most cases—to generate a propulsive force. The quantity, quality, and position of contractions are initiated and coordinated by the nervous system: through this coordination, rhythmic movements of the appendages or body produce locomotion.

      Animals successfully occupy a majority of the vast number of different physical environments (ecological niches) on Earth; in a discussion of locomotion, however, these environments can be divided into four types: aerial (including arboreal), aquatic, fossorial (underground), and terrestrial. The physical restraints to movement—gravity and drag—are the same in each environment: they differ only in degree. gravity is here considered as the weight and inertia (resistance to motion) of a body, drag as any force reducing movement. Although these are not the definitions of a physicist, they are adequate for a general understanding of the forces that impede animal locomotion.

      To counteract the force of gravity, which is particularly important in aerial, fossorial, and terrestrial locomotion, all animals that live in these three environments have evolved skeletal systems to support their body and to prevent the body from collapsing upon itself. The skeletal system may be internal or external, and it may act either as a rigid framework or as a flexible hydraulic (fluid) support.

      To initiate movement, a sufficient amount of muscular work must be performed by aerial, fossorial, and terrestrial animals to overcome inertia. Aquatic (aquatic locomotion) animals must also overcome inertia; the buoyancy of water, however, reduces the influence of gravity on movement. Actually, because many aquatic animals are weightless—i.e., they possess neutral buoyancy by displacing a volume of water that is equal in weight to their dry weight—little muscular work is needed to overcome inertia. But not all aquatic animals are weightless. Those with negative buoyancy sink as a result of their weight; hence, the greater their weight, the more muscular energy they must expend to remain at a given level. Conversely, an animal with positive buoyancy floats to and rests on the surface and must expend muscular energy to remain submerged.

      In water, the primary force that retards or resists forward movement is drag, the amount of which depends upon the animal's shape and how that shape cleaves the water. Drag results mainly from the friction of the water as it flows over the surface of the animal and the adherence of the water to the animal's surface (i.e., the viscosity of the water). Because of the water's viscosity, its flow tends to be lamellar; i.e., different layers of the water flow at different speeds, with the slowest layer of flow being the one adjacent to the body surface. As the flow speed increases, the lamellar pattern is lost, and turbulence develops, thereby increasing the drag.

      Another component of drag is the retardation of forward movement by the backward pull of the eddies of water behind the tail of the animal. As they flow off an animal, the layers of water from each side meet and blend. If the animal is streamlined (e.g., has a fusiform shape), the turbulence is low; if, however, the water layers from the sides meet abruptly and with different speeds, the turbulence is high, causing a strong backward pull, or drag, on the animal.

      Aerial locomotion also encounters resistance from drag, but, because the viscosity and density of air are much less than those of water, drag is also less. The lamellar flow of air across the wing surfaces is, however, extremely important. The upward force of flight, or lift, results from air flowing faster across the upper surface than across the lower surface of the wing. Because this differential in flow produces a lower air pressure on the upper surface, the animal rises. Lift is also produced by the flow of water across surfaces, but aquatic animals use the lift as a steering aid rather than as a source of propulsion.

      Drag is generally considered a negligible influence in terrestrial locomotion; and, in fossorial locomotion, the friction and compactness (friability) of soils are the two major restraints. If the soil is extremely friable, as is sand, some animals can “swim” through it. Such fossorial locomotion, however, is quite rare; most fossorial animals must laboriously tunnel through the soil and thereafter depend upon the tunnels for active locomotion.

      Movement in animals is achieved by two types of locomotion, axial and appendicular. In axial locomotion, which includes the hydraulic ramjet method of ejecting water (e.g., squid), production of a body wave (eel), or the contract–anchor–extend method (leech), the body shape is modified, and the interaction of the entire body with the surrounding environment provides the propulsive force. In appendicular locomotion, special body appendages interact with the environment to produce the propulsive force.

      There are also many animal species that depend on their environment for transportation, a type of mobility called passive locomotion. Some jellyfish, for example, have structures called floats that extend above the water's surface and act as sails. A few spiders have developed an elaborate means of kiting; when a strand of their web silk reaches a certain length after being extended into the air, the wind resistance of the strand is sufficient to carry it away with the attached spider. In one fish, the remora, the dorsal fin has moved to the top of the head and become modified into a sucker; by attaching itself to a larger fish, the remora is able to ride to its next meal.

Microorganisms
      Most motile protozoans (protozoan), which are strictly aquatic animals, move by locomotion involving one of three types of appendages: flagella (flagellum), cilia (cilium), or pseudopodia. Cilia and flagella are indistinguishable in that both are flexible filamentous structures containing two central fibrils (very small fibres) surrounded by a ring of nine double fibrils. The peripheral fibrils seem to be the contractile units and the central ones, neuromotor (nervelike) units. Generally, cilia are short and flagella long, although the size ranges of each overlap.

Flagellar locomotion
      Most flagellate protozoans possess either one or two flagella extending from the anterior (front) end of the body. Some protozoans, however, have several flagella that may be scattered over the entire body; in such cases, the flagella usually are fused into distinctly separate clusters. Flagellar movement, or locomotion, occurs as either planar waves (wave motion), oarlike beating, or three-dimensional waves. All three of these forms of flagellar locomotion consist of contraction waves that pass either from the base to the tip of the flagellum or in the reverse direction to produce forward or backward movement. The planar waves, which occur along a single plane and are similar to a sinusoid (S-shaped) wave form, tend to be asymmetrical; there is a gradual increase in amplitude (peak of the wave) as the wave passes to the tip of the flagellum. In planar locomotion the motion of the flagella is equivalent to that of the body of an eel as it swims. Although symmetrical planar waves have been observed, they apparently are abnormal, because the locomotion they produce is erratic. Planar waves cause the protozoan to rotate on its longitudinal axis, the path of movement tends to be helical (a spiral), and the direction of movement is opposite the propagation direction of the wave.

      In oarlike flagellar movements, which are also planar, the waves tend to be highly asymmetrical, of greater side to side swing, and the protozoan usually rotates and moves with the flagellum at the forward end. In the three-dimensional wave form of flagellar movement, the motion of the flagella is similar to that of an airplane propeller; i.e., the flagella lash from side to side. The flagellum rotates in a conical configuration, the apex (tip) of which centres on the point at which the flagellum is attached to the body. Simultaneous with the conical rotation, asymmetrical sinusoidal waves pass from the base to the end of the flagellum. As a result of the flagellar rotation and its changing angle of contact, water is forced backward over the protozoan, which also tends to rotate, and the organism moves forward in the direction of the flagellum.

Ciliary locomotion
      Cilia operate like flexible oars; they have a unilateral (one-sided) beat lying in a single plane. As a cilium moves backward, it is relatively rigid; upon recovery, however, the cilium becomes flexible, and its tip appears to be dragged forward along the body. Because the cilia either completely cover, as in ciliate protozoans, or are arranged in bands or clumps, the movement of each cilium must be closely coordinated with the movements of all other cilia. This coordination is achieved by metachronal rhythm, in which a wave of simultaneously beating groups of cilia moves from the anterior to the posterior end of the organism. In addition to avoiding interference between adjacent cilia, the metachronal wave also produces continuous forward locomotion because there are always groups of cilia beating backward. Moreover, because the plane of the ciliary beat is diagonal to the longitudinal axis of the body, ciliate organisms rotate during locomotion.

Pseudopodial locomotion
 Although ciliar and flagellar locomotion are clearly forms of appendicular locomotion, pseudopodial locomotion (Figure 1—>) can be classed as either axial or appendicular, depending upon the definition of the pseudopodium. Outwardly, pseudopodial locomotion appears to be the extension of a part of the body that anchors itself and then pulls the remainder of the body forward. Internally, however, the movement is quite different. The amoeba, a protozoan, may be taken as an example. Its cytoplasm (the living substance surrounding the nucleus) is divided into two parts: a peripheral layer, or ectoplasm, of gel (a semisolid, jellylike substance) enclosing an inner mass, or endoplasm, of sol (a fluid containing suspended particles; i.e., a colloid). As a pseudopodium, part of the ectoplasmic gel is converted to sol, whereupon endoplasm begins flowing toward this area, the cell wall expands, and the pseudopodium is extended forward. When the endoplasm, which continues to flow into the pseudopodium, reaches the tip, it extends laterally and is transformed to a gel. Basically, the movement is one of extending an appendage and then emptying the body into the appendage, thereby converting the latter into the former. Although the flow of the cytoplasm is produced by the same proteins involved in the mechanism of muscle contraction, the actual molecular basis of the mechanism is not yet known. Even the mechanics of pseudopodial formation are not completely understood.

Undulating and gliding locomotion
      Two other types of locomotion are observed occasionally in protozoans. Some protozoans, usually flagellates, have along their bodies a longitudinal membrane that undulates, thereby producing a slow forward locomotion. A gliding locomotion is commonly seen in some sporozoans (apicomplexan) (parasitic protozoans), in which the organism glides forward with no change in form and no apparent contractions of the body. Initially, the movement was thought to be produced by ejecting mucus, a slimy secretion; small contractile fibrils have been found that produce minute contraction waves that move the animal forward.

Invertebrates (invertebrate)
      As in the protozoans, aquatic locomotion in invertebrates (animals without backbones) consists of both swimming and bottom (benthos) movements. In swimming, the propulsive force is derived entirely from the interaction between the organism and the water; in bottom movements, the bottom surface provides the interacting surface. Whereas some bottom movements are identical with terrestrial locomotor patterns, others can occur effectively only in the water, where buoyancy is necessary to reduce body weight.

Bottom locomotion
      Small flatworms (flatworm) (Platyhelminthes) and some of the smaller molluscan (mollusk) species move along the bottom by ciliary activity. On their ventral (bottom) surface, a dense coat of cilia extends from head to tail. The direction of the ciliary beat is tailward, causing the animal to glide slowly forward. Generally, all animals that move by this type of ciliary activity secrete a copious stream of mucus over which the animal glides. The mucus not only attaches the animal to the surface but also raises its body so that the cilia can beat. Because ciliary forces are too weak for the movement of large flatworms, they must use muscular contraction for their propulsive force.

      Aquatic invertebrates possess several other types of bottom locomotion. In species with well-developed legs, such as crabs and lobsters, bottom walking is common. Whereas the gaits in such cases are identical to those used on land, they tend to be slightly faster in water, because the buoyancy increases the animal's stability. (Walking gaits are described below; see Terrestrial locomotion (locomotion).)

      Another form of bottom locomotion is bottom creeping, which employs the contract–anchor–extend method of movement. Bottom creeping is best developed in leeches (leech), which have two suckers, one at the anterior end and one at the posterior end. After the posterior sucker anchors the animal, it stretches its body forward and attaches the anterior one. It then releases the posterior sucker and contracts its body toward the anterior end. For effective contract–anchor–extend locomotion, the body musculature must consist of both circular and longitudinal muscles: the contraction of the circular muscles extends or elongates the body; the contraction of the longitudinal muscles flexes and shortens the body. Moreover, the skeleton should be hydrostatic; that is, a fluid skeleton that changes shape but not volume, thereby providing a firm but flexible base.

      In pedal locomotion, which is a slow, continuous gliding that is superficially indistinguishable from ciliary locomotion, propulsion along the bottom is generated by the passage of contraction waves through the ventral musculature, which is in contact with the bottom surface. The pedal contraction waves are either direct (in the same direction as the movement) or retrograde (in the direction opposite to the movement). The direct waves produce locomotion in a manner analogous to that in which a caterpillar walks. When a direct wave reaches a muscle, the muscle contracts and lifts a small part of the body; the body is carried forward and set down anterior to its original position as the wave passes. With direct waves, the surfaces of the body touching the bottom surface are not the ones that contract; with retrograde waves, however, these are the surfaces that do contract. As the retrograde wave approaches, the body area immediately adjacent to it is extended upward. The body surface within the contraction area then anchors itself to the bottom surface, after which the body is pulled forward.

      Large flatworms use pedal locomotion instead of or in alternation with ciliary activity. In the gastropod and amphineuran molluscans (e.g., snails and chitons (chiton), respectively), pedal locomotion is the primary locomotor mode and has become highly complex. The foot of these creeping animals is extremely muscular, penetrated by nerves, and capable of generating one, two, or four laterally adjacent contraction waves. If the foot generates a pair of waves, the lateral halves of the foot may alternate, thereby producing a shuffling movement, or they may be opposite. Generally, a foot can contain no more than one whole and two partial waves moving along a single axis.

      Peristaltic locomotion is a common locomotor pattern in elongated, soft-bodied invertebrates, particularly in segmented worms, such as earthworms (earthworm). It involves the alternation of circular- and longitudinal-muscle-contraction waves. Forward movement is produced by contraction of the circular muscles, which extends or elongates the body; contraction of the longitudinal muscles shortens and anchors the body (see below Fossorial locomotion (locomotion)).

      Although peristaltic locomotion is frequently used by sea cucumbers, they and other echinoderms (echinoderm), such as sea urchins and starfishes (sea star), possess rows of tube feet that provide the main locomotor force. In starfishes, each arm bears hundreds of tube feet. Only one arm, however, becomes dominant in locomotion; while the tube feet on that arm move toward the tip of the arm, the tube feet of the other arms move in the same plane as those of the lead arm. Because there is no apparent metachronal wave of contraction within an arm, the movement of the tube feet is poorly coordinated, but small areas of the tube feet do move in synchrony. Each tube foot is a hollow elastic cylinder capped by a hollow muscular ampulla (a small, bladder-like enlargement). When the ampulla contracts, it forces fluid into the tube foot and extends it. Preferential contraction of muscles in the wall of the tube foot controls the direction of and the retraction of the tube foot. When the tube foot is fully contracted, fluid is withdrawn from it by relaxation of the ampulla, after which the muscles of the tube foot swing it forward in preparation for another step.

      Invertebrates have developed two distinct propulsive mechanisms for swimming: some use hydraulic propulsion; all others utilize undulations of all or parts of their bodies. The medusa (umbrella-shaped) body of coelenterates (cnidarian) and ctenophores (ctenophore) (e.g., jellyfish and comb jelly, respectively) is a flexible hemisphere with tentacles and sense organs suspended from the edge; a manubrium (handle-shaped structure) bearing the digestive system hangs from the internal tip of the hemisphere. Enclosed in the outer margin of the medusa is a wide muscular band; when this band contracts, the opening of the medusa narrows. Simultaneously, water is ejected from the medusa through the narrow opening, and the animal is propelled upward. Because the contractions tend to be regular but slow, locomotion is somewhat jerky.

      Scallops (scallop) are the best swimmers among bivalve molluscans that can swim. Locomotion is produced by rapid clapping movements of the two shells, creating a water jet that propels the scallop. The muscular mantle (a membranous fold beneath the shell) acts as a valve and controls the direction of flow of the ejected water, thereby controlling the direction of movement. Normally, the flow is directed downward on each side of the hinge that joins the two shells, and the resulting water jet lifts the scallop and moves it in the direction of the shell's opening. If necessary, however, escape movement may occur in the opposite direction. The scallop is adapted to swim even though it is two or three times as dense as seawater. The hinge is elastic and opens the shell rapidly; this action, coupled with rapid and repeated contractions of the adductor muscle, which closes the shell, produces a powerful and nearly continuous water jet. Moreover, the body form of a closed scallop is an airfoil (like a wing, the curvature of its upper surface is greater than that of its lower surface); this shape, combined with the downward ejection of water, produces lift.

      Cephalopods (cephalopod) (e.g., squids, octopuses) are another group of mollusks that use hydraulic propulsion. Unlike the scallops, they have lost most of their heavy shell and have developed fusiform bodies. The mantle of cephalopods encloses a cavity in which are contained the gills and other internal organs. It also includes, on its ventral surface, a narrow, funnel-shaped opening (siphon) through which water can be forcibly ejected when all the circular muscles surrounding the mantle cavity contract rapidly and simultaneously. This water jet shoots the cephalopod in a direction opposite to that in which the siphon is pointed.

      Many invertebrates, particularly elongated ones such as open-sea-dwelling annelids and mollusks, swim by undulatory movements produced by contraction waves that alternate on each side of the body. Although the arrangement of the musculature differs between invertebrates and vertebrates, the mechanics of undulatory swimming are the same in both and are described in the following section.

fish and fishlike vertebrates
      Undulatory swimming is roughly analogous to using one oar at the stern of a boat. The side-to-side movements of the oar force the water backward and the boat forward. The undulatory movement of a fish acts similarly, although the motions involved are much more complex.

Anguilliform locomotion
      When an elongated fish such as an eel swims, its entire body, which is flexible throughout its complete length, moves in a series of sinuous waves passing from head to tail. In this type of movement, which is called anguilliform (eel-like) locomotion, the waves cause each segment of the body to oscillate laterally across the axis of movement. Unlike the simple side-to-side movement of the oar, however, each oscillating segment describes a figure-eight loop, the centre of which is along the axis of locomotion. It is these oscillations and the associated orientation of each body segment that produce the propulsive thrust.

      The undulatory body waves are created by metachronal contraction waves alternating between the right and left axial musculature. During steady swimming, several contraction waves simultaneously pass down the body axis from head to tail; the resultant undulatory waves move backward along the body faster than the body moves forward. As the undulatory wave passes backward, its amplitude and speed increase, thereby producing the greatest propulsive thrust in the tail (caudal) region. Propulsion, however, is not limited to the caudal region, for all undulating segments contribute to the thrust. Because the speed, amplitude, and inclination of each body segment differ, the thrust of each differs. In all segments, the greatest thrust is obtained as the segment crosses the locomotor axis, for here it is travelling at its greatest speed and inclination.

Carangiform and ostraciiform locomotion
      All undulatory swimming movements generate forward thrust in the manner described above. Not all swimming animals, however, possess the elongated shape of an eel; only those with a similar body form, in which the surface area of the head end is the same as that of the tail end, have anguilliform locomotion. Fish with fusiform bodies exhibit carangiform locomotion, in which only the posterior half of the body flexes with the passage of contraction waves. This arrangement of body form and locomotion apparently is the most efficient one, for it occurs in the most active and fastest of fish. The advantage of carangiform locomotion appears to be related to the effectiveness of the posterior half of the body as a propulsive unit and the fact that the shape of the body and its small lateral displacement create little water turbulence. In contrast to ostraciiform locomotion, in which only the caudal fin oscillates from side to side in a manner similar to moving a boat with one oar, the length of the propulsive unit of carangiform fish enables the unit to obtain maximum oscillatory speed and inclination.

      Whales (whale) also use undulatory body waves, but unlike any of the fishes, the waves pass dorsoventrally (from top to bottom) and not from side to side. In fact, many mammals that swim mainly by limb movements tend to flex their body in a dorsoventral plane. Whereas the body musculature of fish and tail musculature of amphibians (amphibian) and reptiles (reptile) is highly segmental—that is, a muscle segment alternates with each vertebra—an arrangement that permits the smooth passage of undulatory waves along the body, mammals are unable to produce lateral undulations because they do not have this arrangement. Nor does the muscle arrangement of mammals permit true dorsoventral undulations; however, with an elongated caudal region, as in whales, they can attain a form of carangiform locomotion as effective as that of any fish.

Stabilization and steering
      To stabilize and steer, most aquatic vertebrates have, in addition to the caudal fin, a large dorsal fin and a pair of large anterolateral fins. Although they may possess other fins, these are of less importance. The balance of a swimming animal may be maintained in several ways. Rolling, or rotation, along the longitudinal axis of the body is reduced or controlled by any fins that extend at right angles to the body. Pitching, or dorsoventral seesawing, movements are counteracted by the anterolateral fins, which are also the major steering organs of fish, whales, and seals. Yawing, or lateral seesawing, is prevented by the dorsal fin and, if present, a ventral fin; for these fins to be effective, however, most of their exposed surface area should be behind the fish's centre of gravity. Because fins of the above type are not common in most invertebrates that swim by undulation, their locomotion is less stable.

Tetrapodal vertebrates
      Many of the various types of undulatory locomotion described above are also widely used by aquatic tetrapods (those with walking appendages). Larval frogs, crocodilians, aquatic salamanders, and lizards, for example, have long muscular tails that propel them by undulatory motion. Most aquatic tetrapods, however, move by appendicular locomotion, for which the major propulsive units are the hind legs. The exceptions are sea turtles, auks, penguins (penguin), and fur seals; in these, the hind feet are webbed and are used as rudders. For propulsion, these animals use their forelegs, which have become bladelike flippers in which the forearm and hand region are dorsoventrally compressed to form a single, inflexible unit. The movements of such flippers are analogous to the aerial flight of birds; by moving synchronously, they provide lift and thrust in the water. Unlike aerial flight, however, the upper arms do not produce lift or thrust; instead, they serve only as a pivotal or leverage point for driving the flippers.

      Swimming movements in sea turtles, penguins, and auks (auk) are accomplished by the rotation of the flippers or wings through a figure-eight configuration. In the birds, however, the stroke is relatively faster than in sea turtles, because the entire cycle appears to be proportionately smaller in amplitude. Moreover, because the birds' bodies are more streamlined, they can attain greater speeds than the turtles. Penguins may attain speeds of 40 kilometres (25 miles) per hour in water and have sufficient speed and thrust to enable them to leap two metres (six feet) or more above the water. The wings of penguins are so highly modified, however, that they have lost the ability to fly. The auks, on the other hand, are able to use their wings for both aerial and aquatic locomotion.

      Some of the other aquatic birds, such as ducks (duck) and water ouzels (dipper), are said to propel themselves underwater through wing movements, but the evidence for such propulsion is incomplete and still open to question. The wing movements of ducks may be for steering and hydroplaning (skimming through the water) rather than for actual propulsion. The wings of the water ouzels, or dippers, were once thought to function as hydroplanes, but investigations have revealed that, although the wings are flapped underwater, the ability of dippers to bottom walk or fly underwater depends upon the velocity of the water flowing past the wings rather than the movement of the wings themselves.

      Most aquatic birds are propelled by their webbed hind feet, which tend to move alternately in surface swimming and in unison when the bird is submerged. Of all the swimming birds that use their hind feet, the loons (loon) show the most extreme adaptations: the body, head, and neck are elongated and streamlined; the hind legs are at the very posterior end of the body; the lower legs are compressed and bladelike; and the feet are strongly webbed. The webbing increases the surface area exposed to the water during limb retraction and also permits the folding of the foot, thereby reducing water resistance during protraction.

      In frogs (frog) and freshwater turtles (turtle), the hind legs are elongated and the feet enlarged and strongly webbed. But, whereas the hind legs of frogs move synchronously, except occasionally in slow swimming, when they alternate, the limb movements always alternate in freshwater turtles. Some aquatic turtles, however, such as snapping, mud, and musk turtles, are very poor swimmers and will swim only under extreme conditions. These turtles are bottom walkers, and their limb movements in water are identical to those on land except that they can move faster in water than on land.

      The swimming movements of many mammals are also identical with their terrestrial limb movements. Hippopotamuses spend much of their time in the water, yet they bottom walk rather than swim. Most of the aquatic mammals—e.g., otters, hair seals, aquatic marsupials, insectivores, and rodents—use their hind legs and frequently their tails for swimming. The feet are webbed and usually move alternately; the tail tends to be flattened. Fur seals, polar bears, and platypuses swim mainly with forelimbs; only in the seals, however, are the movements of the forelimbs similar to those of sea turtles and penguins.

Fossorial locomotion (burrowing)
      The speed, manner, and ease with which animals move depends directly on the compactness of the material and its cohesiveness. Many aquatic animals can swim through semisolid mud or muck suspensions, which lack compactness. Some lizards and snakes that live in an arid environment can swim through friable sand, which is compact but lacks strong cohesiveness. Although these swimming movements can be considered a form of fossorial locomotion, the following discussion considers only locomotor patterns in which most of the activity of the animals involved is confined to tunnels that they leave behind.

Fossorial invertebrates
      Burrowing or boring invertebrates have evolved a number of different locomotor patterns to penetrate soil, wood, and stone, of which soil or mud is the easiest to penetrate. The soft-bodied invertebrates (annelid), such as worms (worm) and sea cucumbers (sea cucumber), burrow either by peristaltic locomotion or by the contract–anchor–extend method. Their hydrostatic, or fluid, skeleton, combined with their circular and longitudinal musculature, permits controlled deformation of their shape, which allows them to squeeze into narrow spaces and then enlarge the spaces, thus creating a burrow or tunnel. Worms with a protrusible proboscis (a tubular extension of the oral region) generally burrow by the contract–anchor–extend method. Contraction of the circular muscles in the posterior half of the body drives the body fluids forward, causing the proboscis to evert (turn outward) and forcing it into the soil. When the proboscis is fully everted, the part of the body (collar) directly behind it dilates and anchors the proboscis in the soil. The entire body is then pulled forward by the longitudinal muscles and reanchored. This pattern produces the very jerky and slow forward progression typical of most fossorial locomotion.

      Peristaltic locomotion, which is generated by the alternation of longitudinal- and circular-muscle-contraction waves flowing from the head to the tail, is similar to the above pattern. Forward progression is more continuous, however, because of the contraction waves. The sites of longitudinal contraction are the anchor points; body extension is by circular contraction. The pattern of movement is initiated by anchoring the anterior end. As the longitudinal contraction wave moves posteriorly, it is slowly replaced by the circular contraction wave. The anterior end slowly and forcefully elongates, driving the tip farther into the surface as the circular contraction wave moves down the body. The tip then begins to dilate and anchor the anterior end as another longitudinal contraction wave develops. This sequence is repeated, and the worm moves forward. Reversing the direction of the contraction waves enables the worm to back up.

      Burrowing bivalve mollusks, such as clams (clam), use the contract–anchor–extend locomotor mode. Such bivalves have a large muscular foot that contains longitudinal and transverse muscles as well as a hemocoel (blood cavity). The digging cycle begins with the extension of the foot by contraction of the transverse muscles. The siphons (tubular-shaped organs that carry water to and from the gills) are closed, and the adductor muscle of the shell contracts, thereby forcing blood into the tip of the foot and causing it to dilate. With the tip acting as an anchor, the longitudinal muscles then contract, pulling the body down to the anchored foot. Frequently, the longitudinal muscles contract in short steps and alternate between the left and right sides; this causes the shell to wobble and penetrate deeper as it is pulled down.

      Some invertebrates are able to bore through rock. Most of the rock borers are mollusks; they bore either mechanically by scraping or chemically by the secretion of acid. The piddock, or angel's wing, bivalves, for example, attach themselves to a rock with a sucker-like foot. The two valves, held against the rock, grind back and forth by the alternate contraction of two adductor muscles; the grinding slowly produces a tunnel.

Fossorial vertebrates
      The fossorial vertebrates are found in three classes: amphibians, reptiles, and mammals. Although some fishes and birds dig or bore shallow burrows, they can hardly be considered truly fossorial, as are moles or earthworms. Locomotion of fossorial amphibians and reptiles tends to be axial; it is appendicular only in mammals. Fossorial mammals have strong forelegs with a tendency toward flattening; their hands and particularly the claws are enlarged. Forelegs show the greatest modification in such species as moles (mole) and gophers (pocket gopher), whose entire lives are spent in burrows. These animals tend to dig with a breast stroke, either synchronously or alternately, by extending the foreleg straight forward in front of the snout and then retracting it in a lateral arc. The loosened soil is compacted against the side walls of the burrow. In those fossorial species that dig burrows as nests but forage above the ground—many rodents, such as prairie-dogs, ground squirrels, and groundhogs—the digging movements tend to be dorsoventral with alternating limb movement. The forelegs are extended forward and then retracted downward and backward; the loosened soil passes beneath the body and is frequently pushed to the surface.

      Fossorial reptiles and amphibians are usually legless, or the legs are so reduced that they serve no locomotor function; in most species, the head is flattened dorsoventrally, and the snout extends beyond and somewhat over the mouth. Burrowing is accomplished by one of three patterns analogous to the contract–anchor–extend locomotion of invertebrates. In the most common of these, the snout is driven straight forward along the bottom of the tunnel, the head is then raised, and the soil is compacted to the roof. The head tends to be laterally compressed in animals that use the other two patterns. In one of these patterns, the snout is shoved forward and then swung from side to side; in the other, the snout is rotated as it swings from side to side and seems to shave the walls of the tunnel.

Walking and running
      Only arthropods (arthropod) (e.g., insects (insect), spiders, and crustaceans) and vertebrates have developed a means of rapid surface locomotion. In both groups, the body is raised above the ground and moved forward by means of a series of jointed appendages, the legs. Because the legs provide support as well as propulsion, the sequences of their movements must be adjusted to maintain the body's centre of gravity within a zone of support; if the centre of gravity is outside this zone, the animal loses its balance and falls. It is the necessity to maintain stability that determines the functional sequences of limb movements, which are similar in vertebrates and arthropods. The apparent differences in the walking and slow running gaits of these two groups are caused by differences in the tetrapodal (four-legged) sequences of vertebrates and in the hexapodal (six-legged) or more sequences of arthropods. Although many legs increase stability during locomotion, they also appear to reduce the maximum speed of locomotion. Whereas the fastest vertebrate gaits are asymmetrical, arthropods cannot have asymmetrical gaits, because the movements of the legs would interfere with each other.

Cycle of limb movements
      The cycle of limb movements is the same in both arthropods and vertebrates. During the propulsive, or retractive, stage, which begins with footfall (foot) and ends with foot liftoff, the foot and leg remain essentially stationary as the body pivots forward over the leg. During the recovery, or protractive, stage, which begins with foot liftoff and ends with footfall, the body remains essentially stationary as the leg moves forward. The advance of one leg is a step; a stride is composed of as many steps as there are legs. During a stride, each leg passes through one complete cycle of retraction and protraction, and the distance that the body travels is equal to the longest step in the stride. The speed of locomotion is the product of stride length and duration of stride. Stride duration is directly related to retraction: the longer the propulsive stage, the more time is required to complete a stride and the slower is the gait. A gait is the sequence of leg movements for a single stride. For walking and slow running, gaits are generally symmetrical—i.e., the footfalls are regularly spaced in time. The gaits of fast-running vertebrates, however, tend to be asymmetrical—i.e., the footfalls are irregularly spaced in time.

      The different gaits of insects are based on the synchrony of leg movements on the left (L) and right (R) sides of the animal. The wave of limb movement for each side passes anteriorly; the posterior leg protracts first, then the middle leg, and finally the anterior leg, producing the sequence R3 R2 R1 or L3 L2 L1. There is no limb interference, because the legs of one side do not have footfalls along the same longitudinal axis. The slowest walking gait of insects is the sequence R3 R2 R1 followed by the sequence L3 L2 L1. As the rate of protraction increases, the protractive waves of the right and left sides begin to overlap. Eventually, the top speed is reached when the posterior and anterior legs of one side move synchronously. This gait occurs because the protraction times for all legs are constant, the intervals between posterior and middle legs and between middle and anterior legs are constant, and the interval between posterior and anterior legs decreases with faster movements. Other gaits are possible in addition to those indicated above by altering the synchrony between left and right sides.

      The limb movements of centipedes (centipede) and millipedes (millipede) follow the same general rules as those of insects. The protraction waves usually pass from posterior to anterior. Because each leg is slightly ahead of its anteriorly adjacent leg during the locomotory cycle, one leg touches down or lifts off slightly before its anteriorly adjacent one. This coordination of limb movement produces metachronal waves, the frequency of which equals the duration of the complete protractive and retractive cycle. The length of the wave is directly proportional to the phase lag between adjacent legs.

      Whereas the millipedes must synchronize leg movements to eliminate interference, the tetrapodal vertebrates must synchronize leg movements to obtain maximum stability. Four legs are the minimum requirement for symmetrical terrestrial gaits. Although bipedal (two-legged) gaits require extensive structural modifications of the body and legs, they still retain the leg-movement sequence of tetrapodal gaits. The basic walking pattern of all tetrapodal vertebrates (horse) is left hind leg (LH), left foreleg (LF), right hind leg (RH), right foreleg (RF), and then a cyclic repetition of this sequence, which is equivalent to the slow walking gait of insects but with the middle legs removed. Unlike the insects, however, vertebrates can begin to walk with any of the four legs and not just the posterior pair. The faster symmetrical gaits of vertebrates are obtained by overlapping the leg-movement sequences of the left and right sides in the same manner as insects; for example, an animal can convert a walk to a trot by moving diagonally contralateral legs (those on opposite sides) simultaneously, or to a pace by moving the ipselateral legs (those on the same side) simultaneously. Many other symmetrical gaits occur between the walk and the pace and the trot, which are extreme modifications of the walk.

Cursorial vertebrates
      Cursorial (running) vertebrates are characterized by short, muscular upper legs and thin, elongated lower legs. This adaptation decreases the duration of the retractive–protractive cycle, thereby increasing the animal's speed. Because the leg's cycle is analogous to the swing of a pendulum, reduction of weight at the end of the leg increases its speed of oscillation. Cursorial mammals commonly use either the pace or the trot for steady, slow running. The highest running speeds, such as the gallop, are obtained with asymmetrical gaits. When galloping, the animal is never supported by more than two legs and occasionally is supported by none. The fastest runners, such as cheetahs or greyhounds, have an additional no-contact phase following hind foot contact.

      In cursorial birds and lizards, both of which are bipedal, the feet are enlarged to increase support and the body axis is held perpendicular to the ground, so that the centre of gravity falls between the feet or within the foot-support zone. The running gait is, of course, a simple alternation of left and right legs. In lizards, however, bipedal running must begin with quadrupedal (four-footed) locomotion. As the lizard runs on all four legs, it gradually builds up sufficient speed so that its head end tilts up and back, after which it then runs on only its two hind legs.

Saltation
      The locomotor pattern of saltation (hopping) is confined mainly to kangaroos (marsupial), anurans (tailless amphibians), rabbits (lagomorph), and some groups of rodents in the vertebrates and to a number of insect families in the arthropods. All saltatory animals have hind legs that are approximately twice as long as the anteriormost legs. Although all segments of the hind leg are elongated, two of them—the tibial (between upper segment and ankle) and tarsal (ankle) segments—are the most elongated.

      There are at least four different saltatory patterns, but all are similar in that the simultaneous retraction or extension of the hind legs is followed by an aerial phase of movement. The aerial phase in all patterns is governed by the physical principles of ballistics (the flight characteristics of an object): the height and the length of the jumps are functions of the takeoff velocity and angle. The longest jumps are attained when the takeoff angle is 45°.

      Before jumping, the femur (upper segment of the hind leg) of the flea is held perpendicular to the ground, the tibia extends obliquely posterior, and the remainder of the hind leg extends posteriorly along the ground. Just prior to the jump, the middle legs flex and tilt the body upward; then the femur of the hind legs swings sharply backward simultaneously with the extension of the tibia. This retraction forces the animal upward and forward at an angle of 50°. As the flea approaches touchdown, the front legs are swung forward and downward, the middle legs are held perpendicular to the body axis, and the hind legs project obliquely posterior. The anterior two pairs of legs thus act to absorb the landing shock.

      The frog jump is initiated with three simultaneous movements: the forelegs flex, and the back arches to tilt the entire body upward; the tarsus of the hind leg swings to a vertical position and locks; and the femur, extending anteriorly along the body, swings in a horizontal plane. When the femur is perpendicular to the body, the knee joint snaps open, and the frog jumps forward at a 30° to 45° angle. As the frog begins to land, the forelegs are protracted and held downward in front of the chest. The forefeet touch down first, the forelegs acting as shock absorbers. Simultaneously, the hind legs are protracted so that they can be in jumping posture by the completion of landing.

      The positions and movements of the hind legs in rabbits and kangaroos are similar to those of the frog. The major difference is that rabbits, kangaroos, and all other mammals move their legs in a vertical plane instead of a horizontal plane, as do the frogs; because the femur and tibia move vertically, the tarsus need not be elevated to prevent the hind leg from hitting the ground. The saltatorial gait of rabbits is quadrupedal, whereas that of kangaroos is bipedal. A jumping rabbit stretches forward and lands on its forefeet; generally, both forefeet do not touch ground simultaneously, however. As the forefeet touch, the back flexes, and the hind end rotates forward and downward. The hind feet touch down lateral to the forefeet, and, as the back extends, a new jump begins. In contrast, the kangaroo lands on its hind feet, and the back is held fairly straight through all phases of the jump, although the body inclines forward at takeoff and posteriorly when landing.

Crawling
      Invertebrates crawl either by peristaltic locomotion or by contract–anchor–extend locomotion, both of which have been described previously (see above Fossorial locomotion (locomotion)). Limbless vertebrates, however, crawl in one of four patterns: serpentine, rectilinear, concertina, and sidewinding. The most common pattern, serpentine locomotion, is used by snakes (snake), legless lizards, amphisbaenids (worm lizards), and caecilians (wormlike amphibians). Rectilinear locomotion is used by the giant snakes and almost exclusively by fossorial vertebrates when burrowing. Concertina and sidewinding locomotion are largely confined to snakes.

Serpentine locomotion
      In serpentine locomotion, in which the body is thrown into a series of sinuous curves, the movements appear identical to those of anguilliform swimming, but the similarity is more apparent than real. Unlike anguilliform swimming, when a snake starts to move, the entire body moves, and all parts follow the same path as the head. When the snake stops moving, the entire body stops simultaneously. Propulsion is not by contraction waves undulating the body but by a simultaneous lateral thrust in all segments of the body in contact with solid projections (raised surfaces). The muscular thrust against the projection is perpendicular to the axis of the pushing segment. To go forward, therefore, it is necessary for the strongest thrust to act against the side of the projection facing in the direction of movement. Because of this, thrust tends to occur at the anterior end of the concave (inward-curving) side of the loop of the snake's body.

Concertina locomotion
      Concertina locomotion is used when there is not enough frictional resistance along the locomotor surface for serpentine locomotion. After the body is thrown into a series of tight, sinuous loops, forming a frictional anchor, the head slowly extends forward until the body is nearly straight or begins to slide. The anterior end forms a small series of loops and, with this anchor, pulls the posterior regions forward, after which the sequence of movements is repeated. This crawling pattern is analogous to the contract–anchor–extend locomotion of invertebrates, but, because snakes lack the body flexibility provided by a hydrostatic skeleton, they must depend upon the body loops.

Sidewinding
      Sidewinding, which is also used when the locomotor surface fails to provide a rigid frictional base, is a specific adaptation for crawling over friable sandy soils. Like serpentine locomotion but unlike concertina locomotion, the entire body of the snake moves forward continuously in sidewinding locomotion. Although the body moves through a series of sinuous curves, the track made by the snake is a set of parallel lines that are roughly perpendicular to the axis of movement. This happens because only two parts of the body touch the ground at any instant; the remainder of the body is held off the ground. To begin sidewinding, the snake arches the anterior part of the body forward and forms an elevated loop with only the head and the middle of the body in contact with the ground. Because each part of the body touches the ground only briefly before it begins to arch forward again, the snake seems to roll forward much like a short, coiled spring. In a continuously repeating cycle, as a segment arches forward, the posteriorly adjacent segment touches down.

Rectilinear locomotion
      Unlike the three preceding patterns of movement, in which the body is thrown into a series of curves, in rectilinear locomotion in snakes the body is held relatively straight and glides forward in a manner analogous to the pedal locomotion of snails. The ventral (belly) surface of snakes is covered by scales (scale) elongated crosswise that overlap like roof shingles, with the opening of the overlap facing toward the posterior. Each ventral scale is moved by two pairs of muscles, both of which are attached to ribs but not to ribs of the same segment as the scale. One pair of muscles is inclined posterior at an angle (obliquely); the other is inclined anterior at an angle. As contraction waves move rearward from the head simultaneously on both sides, the anterior oblique muscles of a scale contract first and lift the scale upward and forward. When the posterior oblique muscles contract, the scale is pulled rearward, but its edge anchors it, and the body is pulled forward. This sequence is repeated by all segments as the contraction wave passes posteriorly, and, as a series of contraction waves follow one another, the body slowly inches forward.

Arboreal and aerial locomotion

Climbing
      The adaptation for climbing is unique for each group of arboreal animals. All climbers must have strong grasping abilities, and they must keep their centre of gravity as close as possible to the object being climbed. Because arthropods are generally small and, thus, not greatly affected by the pull of gravity, they show little specific structural adaptation for climbing. In contrast, the larger and heavier-bodied vertebrates have many climbing specializations. In both arthropods and vertebrates, however, no leg is moved until the others are firmly anchored.

Arboreal amphibians and reptiles
      Arboreal frogs are slender-bodied anurans with tapering legs and feet. The tips of the toes (digits) are expanded into large, circular disks that may function as suction cups, although such an action has not yet been definitely demonstrated. The disks, however, do increase the contact area, thereby improving grasping ability. The leg-movement sequence during climbing is that of a walking gait.

      Arboreal lizards have the same type of climbing gait as arboreal frogs, and their climbing specializations are also similar to those of anurans. They have a different type of climbing foot, however, because of the presence of claws and scales on the digits. Moreover, the entire digits, rather than just their tips, may be expanded. On the bottom of each of these spatula-shaped expansions are one or two rows of transversely elongated scales. Although not visible to the naked eye, the surface of these scales is covered with fine projections that increase their ability to adhere to a surface. Because of this strong adherence, the toes roll off and on the surface on which the animal is walking. Unlike other arboreal lizards, chameleons (chameleon) possess a prehensile (grasping) tail and zygodactylous feet—i.e., the toes are fused into two opposable units. Although these adaptations are inferior for vertical climbing, they are superior for locomotion on vertical or inclined, slender branches. Arboreal snakes tend to have either prehensile tails or extremely elongated bodies.

Climbing birds and mammals
      Although the strong, clawed feet of birds permit many of them to climb occasionally, most truly scansorial (climbing) birds cling with their strong feet and brace themselves with stiffened tail feathers. Birds such as woodpeckers and tree creepers usually climb vertically upward, usually with both feet moving simultaneously in short, vertical hops. This mode of locomotion, however, prevents vertical descent. Only the nuthatch can descend as easily as it can ascend; it climbs obliquely, using the upper foot for clinging and the lower foot as a brace. Parrots (parrot) have developed zygodactylous feet as an aid to climbing; in addition, they frequently use their bills when climbing vertically.

      Several locomotor patterns for climbing are used by arboreal mammals, the grasping ability of which has been enhanced by the presence of either strong claws or prehensile fingers. Many monkeys (monkey) use a climbing gait similar to the leg sequence of walking. Occasionally, however, they use a leg sequence equivalent to that of a trot. Small-bodied climbers with sharp claws, such as squirrels (squirrel), climb by the alternate use of forelegs and hind legs; essentially, they hop up a tree. Prehensile-fingered climbers descend backward and generally with a walking type of leg sequence. Sharp-clawed species descend with a similar gait sequence but with the head downward.

Leaping
      The mechanics of arboreal leaping do not differ from those of terrestrial saltation; the upward thrust in both is produced by the rapid, simultaneous extension of the hind legs. Because of the narrowness of the arboreal landing site, however, landing behaviour does differ. Arboreal leaping also tends to be a discontinuous locomotor behaviour that is used only to cross wide gaps in the locomotor surface. Leaping from limb to limb, although occasionally employed by most climbers, appears to occur most frequently in animals with opposable or at least prehensile forefeet, particularly tree frogs and primates. Such forefeet enable the animal to grasp and hold onto the landing site.

      True brachiation (using the arms to swing from one place to another) is confined to a few species of primates, such as gibbons and spider monkeys. Because the body is suspended from a branch by the arms, brachiation is strictly foreleg locomotion. When the animal moves, it relaxes the grip of one hand, and the body pivots on the shoulder of the opposite arm and swings forward; then the free arm reaches forward at the end of the body's swing and grabs a branch. The sequence is then repeated for the other arm. This locomotor pattern produces a relatively rapid and continuous forward movement but is restricted to areas with thick canopies of trees. Brachiators have arms that may be as long or longer than the body and a very motile shoulder joint.

Gliding
      There are two functionally distinct forms of gliding, gravitational gliding and soaring: the former is used by gliding amphibians, reptiles, and mammals; the latter is restricted to birds. All gliders are able to increase the relative width of their bodies, thereby increasing the surface area exposed to wind resistance. The few gliding frogs flatten their bodies dorsoventrally and spread their limbs outward. Gliding snakes not only flatten their bodies but also draw in the ventral scales, thereby creating a trough. The best-adapted gliding lizards have elongated ribs that open laterally like a fan.

      Gliding mammals, such as the African flying squirrel and the colugo (flying lemur), usually have, on each side of the body, a fold of skin (the patagium) that extends from their wrist or forearm backward along the body to the shank of the hind leg or the ankle. When gliding, they assume a spread-eagle posture, and the patagia unfold.

Gravitational gliding
      Gravitational gliding is equivalent to parachuting. Because the expanded lateral surface of the body increases the wind resistance against the body, the speed of falling is reduced. The directions of gliding can be controlled by adjusting the surface area—to curve to the right, the right patagium is relaxed. Gliders can land on vertical surfaces by suddenly turning the anterior end of the body up as it reaches the surface. Mechanically, this stalls the flight—i.e., the horizontal component of flight is eliminated.

      Gravitational gliding is one of the basic mechanisms of soaring, which is restricted to birds, although birds must obtain their initial elevation by means of flapping flight. The second basic mechanism of soaring involves wind or air currents. Soaring requires that air currents meet one of two conditions: either the air must have a vertical velocity exceeding the rate of descent in gravitational gliding, or it must have a horizontal velocity that is nonuniform in time and space. Whereas static soaring depends upon vertical air currents, dynamic soaring depends upon horizontal air currents. Both types of soaring are described below.

      Vertical air currents for static soaring are produced when wind strikes an obstruction and is deflected upward. The sites of deflection are very local and discontinuous and seldom extend more than 30 metres (100 feet) above the obstruction. The height of deflection and the vertical velocity of the air are a function of the angle of deflection and the velocity of the wind. If the vertical velocity of the air equals the descent speed of the bird, the bird remains stationary in height relative to the ground. If, however, the vertical velocity is greater, the bird rises, and, if less, the bird falls at a speed equal to the gravitational descent speed minus the air's vertical ascent speed. The horizontal velocity of the air determines the bird's movements relative to the ground in the same manner as that of the vertical velocity.

      The soaring flights of vultures (vulture) and hawks (hawk) depend upon vertical hot-air currents called thermals (thermal). Such currents are not continuous updrafts or downdrafts originating from a specific spot; instead, as a local region of the ground is heated, a vertical, hot-air updraft is created. At the top of the column, a thermal bubble is formed by the hot air curving outward, downward, and then around the bubble. It is then pinched off by cool air flowing into the column and floats upward. The free-floating thermal bubble is doughnut shaped, with the air rising in the centre and cycling outward and downward. Soaring birds spiral downward in the updraft; however, because the bubble rises faster than birds descend, soaring birds are carried upward, but at a speed less than that of the bubble. When a bird reaches the bottom of the bubble, it begins a straight gravitational glide until it reaches the next thermal bubble. Thus, static soaring in a thermal bubble can be recognized by its alternating flight pattern of circling and straight gliding.

      Unlike static soaring, which is done at relatively high altitudes over land, dynamic soaring is done at low levels and is usually restricted to oceanic areas. Dynamic soaring depends upon a steady horizontal sea wind, which is laminated into layers of different velocities because of the frictional interaction between the water and the air; the lower layers have the lowest velocity. The flight path of a bird performing dynamic soaring tends to be a series of inclined loops that are perpendicular to the direction of the wind. A soaring albatross, for example, will begin its gravitational glide approximately 15 metres (50 feet) above the sea. Because it glides downwind, its velocity is increased both by descent and by the wind at its tail. As the bird nears the sea, it makes a turn into the wind, and the forward flight velocity derived from the downwind glide and the tail wind combine to lift the albatross slowly back to its initial gliding height, but with a loss of horizontal velocity. The bird therefore turns downwind again and begins to repeat the soaring cycle.

      Because it depends upon the presence of a horizontal air current, the flight of flying fish is more akin to soaring than to true flying. As a flying fish approaches the water surface, its pectoral and pelvic fins, which are analogous to the forelimbs and hind limbs of quadrupeds, are pressed along the side of the body. The greatly enlarged, winglike pectoral fins then spread out as the fish leaves the water. The wind against the fins provides lift to raise the body above the water, and the tail continues to undulate to provide additional thrust. When the entire body is out of the water, the enlarged pelvic fins extend, and the fish glides for a short distance until its forward velocity is lost. Occasionally, as a fish drops back into the water, it will undulate its tail to initiate another short flight.

True flight
      Three animal groups have developed true flight: insects, birds, and mammals. All generate forward thrust by flapping lateral appendages (wing), and all are free of any dependence on gravitational descent or air currents. It should be noted at the outset, however, that, although the aerodynamics of flight are identical in all three, the following cycles of wing movements described for the different animal groups are generalizations; each species in a group has a distinctive flight pattern and, therefore, a distinctive pattern of wing movement.

      Flight is produced by the simultaneous rotation of the left and right wings in a circle or in a figure eight. This rotation produces the upward thrust, or lift, necessary to overcome gravity and the forward thrust required to overcome drag. As the downward and backward phase of rotation forces the air backward and the body forward, lift is produced by the unequal velocities of the air across the upper and lower wing surfaces.

Wings of insects
      In flies (fly) with one pair of wings, the rotation of the tip inscribes a posterior inclined oval. At the top of the wing cycle, the tip lies above the junction of the thorax and abdomen. The wing then beats downward and forward so that the tip ends anterior and below the head. To insure maximum thrust, the broad surface of the wing lies parallel to the horizontal body plane during the downstroke. During the path of the upstroke, which is the reverse of the downstroke, the wing is feathered (turned) by inclining it perpendicular to the body plane. Although the rotational cycle of those insects with two pairs of wings follows a similar path, the upward and downward strokes of the anterior and posterior wings are not simultaneous; the anterior pair usually lags behind the posterior pair.

      The wings of insects are rotated by pulsation of the thorax, not by a set of muscles. Basically, the thorax is a rigid box to which the wings are attached by a pair of longitudinal lateral hinges that enable the thorax to move dorsoventrally. Four sets of muscles control the major movements. Contraction of a perpendicular set, which extends from the centre of the floor of the thorax to its roof, depresses the thorax and, because of a reverse linkage between wing and thorax, raises the wing. Contraction of a diagonal set, which extends from the anterior roof of the thorax to its posterior floor, elevates the thorax and lowers the wing. Two diagonal sets of muscles extend laterally from the floor to the wall of the thorax and are responsible for maintaining a relatively constant width in the thorax.

Wings of birds and bats
      Unlike insect wings, the wings of birds and bats are linked structures, the lateral extent and regional inclination of which are altered intrinsically by muscular and bony segments. The up-and-down strokes of a bird's wing are produced by large chest (pectoral) muscles that extend from the sternum (breastbone) to the lower surface of the humerus (a bone in the upper arm). When these muscles contract, the wing is lowered; it is raised by the contraction of a small anterior pectoral muscle that is attached to the upper surface of the humerus by a long tendon.

      Birds exhibit two major flight patterns, hovering flight and propulsive flight. Hovering flight is of fairly restricted use and is observed most frequently in the hummingbirds (hummingbird). The path of the wings inscribes a horizontal figure eight whose centre is perpendicular to the shoulder joint. The downward stroke of the wings is actually a slightly inclined anterior stroke, and, because the longitudinal body axis is nearly perpendicular to the ground, the upward stroke is a horizontal posterior stroke. Both strokes are power strokes that produce lift: on the downstroke the dorsal wing surface is the top of the airfoil surface; on the upstroke the ventral surface is the top of the airfoil surface.

      Most birds and bats, however, utilize propulsive flight. Because the body is not stationary, as it is in hovering flight, the wing always moves forward relative to the air, and its tip generally inscribes an oval or figure-eight path. In a pigeon, for example, the downstroke begins with the wing fully extended and perpendicular to the back. As the wing moves downward and anterior, it draws level with the body, at which point the upper arm section stops while the distal part completes the downward path. At the bottom of the downstroke, the distal part turns outward and is elevated rapidly by the combined protraction of the humerus and the extension of the distal section.

Directional control
      Although an animal's locomotor pattern may be controlled by its nervous system (stereotyped response), directional control is impossible without sensory input. Two factors are involved in directional control: orientation, the ability of an animal to determine and to alter its position in the environment; and steering, the mechanical alteration of the locomotor pattern through which the animal adjusts its position.

Orientation
      Orientation of locomotor behaviour is usually categorized as either kinesis or taxis. In kinesis, as previously explained, an animal's body is not oriented in relation to a sensory stimulus; rather, the stimulus causes an alteration in speed or direction of movement. In wood lice, for example, the kinetic response alters only the rate of movement. Because wood lice tend to aggregate in moist areas, their ambulatory activity increases or decreases as the relative humidity decreases or increases, respectively. In the planarian (an aquatic, ciliated flatworm), on the other hand, the kinetic response affects only the rate at which the planarian changes its direction. Because planaria tend to stay in or return to darker areas, an increase in light intensity causes an increase in their turning responses. Generally, however, animals tend to alter both direction and speed as a single kinetic response.

      In taxis, an animal orients itself in a specific spatial relationship to a stimulus. The orientation may be simply an alteration of body position or it may be an alteration of locomotor direction so that the animal moves toward, away from, or at a fixed angle to the source of the stimulus. Sources that elicit a taxis response, which may cause a modification of speed, direction, or both, seem to encompass the entire range of environmental stimuli, such as gravity (geotaxis), temperature (thermotaxis), light (phototaxis), or chemicals (chemotaxis). If the response is negative, the animal moves away from the source; if it is positive, the animal moves toward the source.

      The control of the response to a taxis is of two types. In open-system control, the initial response to a stimulus has no effect on subsequent responses to the same stimulus. A male firefly, for example, locates a female by the latter's brief flashes of light. When a male sees a female's flash, the male turns in the direction of the female, even though the source is no longer visible. If another female flashes, however, the male responds to the second flash in exactly the same manner as it did to the first. In close-system control, on the other hand, the response is progressively altered by feedback so that all subsequent responses are adjusted to the initial response. A bat chasing a flying insect will alter its flight path to intercept that of the insect. The bat's initial change in direction is only a general alteration of its course, but, as it approaches the insect, the bat constantly modifies its course to obtain an accurate interception.

Steering
      Animals obtain accurate directional response (steering) by changing their propulsive response. Because steering relies heavily on continuous feedback (the communication cycle in which the motor output, or behaviour, is constantly being modified by the sensory input, or stimulus), it requires a precise integration of the central and peripheral nervous systems (nervous system). (The central nervous system—in vertebrates, the brain and spinal cord—is that part of the nervous system that receives sensory impulses and sends out motor impulses; the peripheral nervous system consists of all the nerves that carry impulses between the central nervous system and other parts of the body.) Exteroceptive stimuli (those that originate outside the body) received by the peripheral nervous system establish the animal's spatial position in the environment; proprioceptive stimuli (those that originate inside the body), also received by the peripheral nervous system, establish the relative position of the body units to each other. Through integration of these two sets of stimuli, the central nervous system continuously adjusts the contraction of the motor units (e.g., muscles) to obtain the desired orientation.

      During locomotion, steering is a continual process. The direction of movements must be constantly adjusted to counteract environmentally produced deviations of direction. The apparently simple act of a bird flying from a tree to the ground illustrates the complexity of directional control. As the bird flies to the ground, it must be constantly aware of its height above the ground, the orientation of its body axis relative to the ground, deviations in flight direction resulting from air currents, and its speed of fall. All these parameters are determined primarily by exteroceptive stimuli received through the eyes and inner ears. The downward flight is constantly adjusted in response to these exteroceptive stimuli, and the fine control necessary for these adjustments is obtained by proprioceptive feedback.

George R. Zug Ed.

Additional Reading
Eadweard Muybridge, Animals in Motion (1899, reissued 1957), is a classic work on the subject. R.A.R. Tricker and B.J.K. Tricker, The Science of Movement (1966), is a useful introduction to the physics of locomotion. James Gray, Animal Locomotion (1968), provides a synthesis of most aspects of invertebrate and vertebrate locomotion. R.B. Clark, Dynamics in Metazoan Evolution (1964), discusses locomotor patterns of invertebrates with hydrostatic skeletons. Heinrich Hertel, Structure, Form, and Movement (1966; originally published in German, 1963), treats the mechanics of flight and undulatory swimming.

* * *


Universalium. 2010.

Игры ⚽ Поможем решить контрольную работу
Synonyms:
(from place to place), /


Look at other dictionaries:

  • locomotion — [ lɔkɔmosjɔ̃ ] n. f. • 1771; de locomotif, d apr. le lat. motio 1 ♦ Didact. Action de se mouvoir, de se déplacer d un lieu vers un autre; fonction qui assure ce mouvement. Muscles de la locomotion. ⇒ locomoteur. La marche, mode naturel de la… …   Encyclopédie Universelle

  • Locomotion № 1 — Locomotion № 1 …   Википедия

  • Locomotion n°1 — Locomotion n° 1 Pour les articles homonymes, voir Locomotion. Locomotion …   Wikipédia en Français

  • Locomotion — Lo co*mo tion, n. [L. locus place + motio motion: cf. F. locomotion. See {Local}, and {Motion}.] 1. The act of moving from place to place. Animal locomotion. Milton. [1913 Webster] 2. The power of moving from place to place, characteristic of the …   The Collaborative International Dictionary of English

  • Locomotion — (von lat. locus „Ort, Stelle“ und motio „Bewegung“) bezeichnet: ein durch eine Coverversion von Kylie Minogue nochmals bekannt gewordenes Lied von Little Eva (siehe: The Loco Motion), sowie einen Modetanz aus den 1960er Jahren eine… …   Deutsch Wikipedia

  • Locomotion —   [englisch, ləʊkə məʊʃn], geselliger Modetanz; Anfang der Sechzigerjahre in England unter Zugrundelegung einer langsamen, twistähnlichen Musik im 4/4 Takt entstanden. Typisches Beispiel: Little Eva (* 1943) mit »(Do The) Locomotion« (1962) …   Universal-Lexikon

  • Locomotion — es un canal latinoamericano de televisión por cable, que emite 24 horas diarias de dibujos animados. Su sede administrativa se encuentra en Miami (Estados Unidos) y su centro de emisiones en Buenos Aires (Argentina). Creado en 1997, fue… …   Enciclopedia Universal

  • locomotion — 1640s, formed in English from L. loco from a place (ablative of locus place ) + motionem (nom. motio) motion, a moving …   Etymology dictionary

  • locomotion — *motion, movement, move, stir …   New Dictionary of Synonyms

  • locomotion — [n] movement action, mobileness, mobility, motion, moving, progression, travel, travelling; concepts 2,145,697 …   New thesaurus

  • locomotion — ► NOUN ▪ movement or the ability to move from one place to another. DERIVATIVES locomotory adjective. ORIGIN from Latin loco from a place + motio motion …   English terms dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”