photoreception

photoreception
/foh'toh ri sep"sheuhn/, n.
the physiological perception of light.
[1905-10; PHOTO- + RECEPTION]

* * *

Biological responses to stimulation by light, most often referring to the mechanism of vision.

In one-celled organisms such as the amoeba, the whole cell may be sensitive to light. Earthworms have photoreceptive cells scattered over their bodies to help orient themselves by comparison of light intensities in different directions. Most animals have localized photoreceptors of varying complexity. In humans, photoreception relies on the chemical response of a light-sensitive pigment, rhodopsin, in photoreceptor cells in the retina of the eye. Stimulation of those cells results in a stimulus being conducted toward the nervous system. Humans, like other vertebrates, have two types of photosensitive cells, rod cells and cone cells. Rod cells are responsible for vision when there is little light; cone cells mediate daylight vision and colour. Photoreception also refers to photosynthesis in plants. See also sense.

* * *

Introduction
 any of the biological responses of animals (animal) to stimulation by light.

      In animals photoreception refers to mechanisms of light detection that lead to vision and depends on specialized light-sensitive cells (cell) called photoreceptors, which are located in the eye (eye, human). The quality of vision provided by photoreceptors varies enormously among animals. For example, some simple eyes such as those of flatworms (flatworm) have few photoreceptors and are capable of determining only the approximate direction of a light source. In contrast, the human eye has 100 million photoreceptors and can resolve one minute of arc (one-sixtieth of a degree), which is about 4,000 times better than the resolution achieved by the flatworm eye.

      The following article discusses the diversity and evolution of eyes, the structure and function of photoreceptors, and the central processing of visual information in the brain. For more information about the detection of light, see optics; for general aspects concerning the response of organisms to their environments, see sensory reception (senses).

Diversity of eyes
 The eyes of animals are diverse not only in size and shape but also in the ways in which they function. For example, the eyes of fish from the deep sea often show variations on the basic spherical design of the eye. In these fish, the eye's field of view is restricted to the upward direction, presumably because this is the only direction from which there is any light from the surface. This makes the eye tubular in shape. Some fish living in the deep sea have reduced eyelike structures directed downward (e.g., Bathylychnops, which has a second lens and retina attached to the main eye); it is thought that the function of these structures is to detect bioluminescent (bioluminescence) creatures. On the ocean floor, where no light from the sky penetrates, eyes are often reduced or absent. However, in the case of Ipnops, which appears to be eyeless, the retina is still present as a pair of plates covering the front of the top of the head, although there is no lens or any other optical structure. The function of this eye is unknown.

  The placing of the eyes in the head varies. Predators, such as felines (feline) and owls (owl), have forward-pointing eyes and the ability to judge distance by binocular triangulation. Herbivorous species that are likely to be victims of predation, such as mice (mouse) and rabbits (rabbit), usually have their eyes almost opposite each other, giving near-complete coverage of their surroundings. In addition to placement in the head, the structure of the eye varies among animals. Nocturnal animals, such as the house mouse and opossum, have almost spherical lenses filling most of the eye cavity. This design allows the eye to capture the maximum amount of light possible. In contrast, diurnal animals, such as humans and most birds (bird), have smaller, thinner lenses placed well forward in the eye. Nocturnal animals usually have retinas with a preponderance of photoreceptors called rods (rod), which do not detect colour but perceive size, shape, and brightness. Strictly diurnal animals, such as squirrels (squirrel) and many birds, have retinas containing photoreceptors called cones (cone), which perceive both colour and fine detail. A slit pupil is common in nocturnal animals, as it can be closed more effectively in bright light than a round pupil. In addition, nocturnal animals, such as cats and bush babies (bush baby), are usually equipped with a tapetum, a reflector behind the retina designed to give receptors a second chance to catch photons (photon) that were missed on their first passage through the retina.

      Animals such as seals (seal), otters (otter), and diving birds, which move from air to water and back, have evolved uniquely shaped corneas (cornea)—the transparent membrane in front of the eye that separates fluids inside the eye from fluids outside the eye. The cornea functions to increase the focusing power of the eye; however, optical power is greatly reduced when there is fluid on both sides of the membrane. As a result, seals, which have a nearly flat cornea with little optical power in air or water, rely on a re-evolved spherical lens to produce images. Diving ducks (duck), on the other hand, compensate for the loss of optical power in water by squeezing the lens into the bony ring around the iris, forming a high curvature blip on the lens surface, which shortens its focal length (the distance from the retina to the centre of the lens). One of the most interesting examples of amphibious optics occurs in the “four-eyed fish” of the genus Anableps (four-eyed fish), which cruises the surface meniscus with the upper part of the eye looking into air and the lower part looking into water. It makes use of an elliptical lens, with the relatively flat sides adding little to the power of the cornea and the higher curvature ends focusing light from below the surface, where the cornea is ineffective.

      Though the eyes of animals are diverse in structure and use distinct optical mechanisms to achieve resolution, eyes can be differentiated into two primary types: single-chambered and compound. Single-chambered eyes (sometimes called camera eyes) are concave structures in which the photoreceptors are supplied with light that enters the eye through a single lens. In contrast, compound eyes are convex structures in which the photoreceptors are supplied with light that enters the eye through multiple lenses. The possession of multiple lenses is what gives these eyes their characteristic faceted appearance.

Single-chambered eyes
Pigment cup eyes
      In most of the invertebrate phyla, eyes consist of a cup of dark pigment that contains anywhere from a few photoreceptors to a few hundred photoreceptors. In most pigment cup eyes there is no optical system other than the opening, or aperture, through which light enters the cup. This aperture acts as a wide pinhole and restricts the width of the cone of light that reaches any one photoreceptor, thereby providing a very limited degree of resolution. Pigment cup eyes are very small, typically 100 μm (0.004 inch) or less in diameter. They are capable of supplying information about the general direction of light, which is adequate for finding the right part of the environment in which to seek food. However, they are of little value for hunting prey or evading predators. In 1977 Austrian zoologist Luitfried von Salvini-Plawen and American biologist Ernst Mayr (Mayr, Ernst) estimated that pigment cup eyes evolved independently between 40 and 65 times across the animal kingdom. These estimates were based on a variety of differences in microstructure among pigment cup eyes of different organisms. Pigment cup eyes were undoubtedly the starting point for the evolution of the much larger and more optically complex eyes of mollusks (mollusk) and vertebrates (vertebrate).

Pinhole eyes
 Pinhole eyes, in which the size of the pigment aperture is reduced, have better resolution than pigment cup eyes. The most impressive pinhole eyes are found in the mollusk genus nautilus, a member of a cephalopod group that has changed little since the Cambrian Period. These organisms have eyes that are large, about 10 mm (0.39 inch) across, with millions of photoreceptors. They also have muscles (muscle) that move the eyes and pupils that can vary in diameter, from 0.4–2.8 mm (0.02–0.11 inch), with light intensity. These features all suggest an eye that should be comparable in performance to the eyes of other cephalopods, such as the genus octopus. However, because there is no lens and each photoreceptor must cover a wide angle of the field of view, the image in the Nautilus eye is of very poor resolution. Even with the pupil at its smallest, each receptor views an angle of more than two degrees, compared with a few fractions of a degree in Octopus. In addition, because the pupil has to be small in order to achieve even a modest degree of resolution, the image produced in the Nautilus eye is extremely dim. Thus, a limitation of pinhole eyes is that any improvement in resolution is at the expense of sensitivity; this is not true of eyes that contain lenses. There are one or two other eyes in gastropod mollusks that could qualify as pinhole eyes, notably those of the abalone genus Haliotis. However, none of these eyes rival the eyes of Nautilus in size or complexity.

Lens eyes
      Relative to pinhole eyes, lens eyes have greatly improved resolution and image brightness. Lenses (lens) were formed by increasing the refractive index of material in the chamber by adding denser material, such as mucus or protein. This converged incoming rays of light, thereby reducing the angle over which each photoreceptor receives light. The continuation of this process ultimately results in a lens capable of forming an image focused on the retina. Most lenses in aquatic animals are spherical, because this shape gives the shortest focal length for a lens of a given diameter, which in turn gives the brightest image. Lens eyes focus an image either by physically moving the lens toward or away from the retina or by using eye muscles to adjust the shape of the lens.

      For many years the lens properties that allow for the formation of quality images in the eye were poorly understood. Lenses made of homogeneous material (e.g., glass or dry protein) suffer from a defect known as spherical aberration, in which peripheral rays are focused too strongly, resulting in a poor image. In the 19th century, Scottish mathematician and physicist James Clerk Maxwell (Maxwell, James Clerk) discovered that the lens of the eye must contain a gradient of refractive index, with the highest degree of refraction occurring in the centre of the lens. In the late 19th century the physiologist Matthiessen showed that this was true for fish, marine mammals (mammal), and cephalopod mollusks. It is also true of many gastropod mollusks, some marine worms (marine worm) (family Alciopidae), and at least one group of crustaceans (crustacean), the copepod genus Labidocera. Two measurements, focal length and radius of curvature of the lens, can be used to distinguish gradient lenses from homogeneous lenses. For example, gradient lenses have a much shorter focal length than homogeneous lenses with the same central refractive index, and the radius of curvature of a gradient lens is about 2.5 lens radii, compared with 4 radii for a homogeneous lens. The ratio of focal length to radius of curvature is known as the Matthiessen ratio (named for its discoverer, German physicist and zoologist Ludwig Matthiessen) and is used to determine the optical quality of lenses.

      The lens eyes of fish and cephalopod mollusks are superficially very similar. Both are spherical and have a Matthiessen ratio lens that can be focused by moving it toward and away from the retina, an iris that can contract, and external muscles that move the eyes in similar ways. However, fish and cephalopod mollusks evolved quite independently of each other. An obvious difference between the eyes of these organisms is in the structure of the retina. The vertebrate retina is inverse, with the neurons (neuron) emerging from the front of the retina and the nerve fibres burrowing out through the optic disk at the back of the eye to form the optic nerve. The cephalopod retina is everse, meaning the fibres of the neurons leave the eye directly from the rear portions of the photoreceptors. The photoreceptors themselves are different too. Vertebrate photoreceptors, the rods (rod) and cones (cone), are made of disks derived from cilia (cilium), and they hyperpolarize (become more negative) when light strikes them. In contrast, cephalopod photoreceptors are made from arrays of microvilli (fingerlike projections) and depolarize (become less negative) in response to light. The developmental origins of the eyes are also different. Vertebrate eyes come from neural tissue, whereas cephalopod eyes come from epidermal tissue. This is a classic case of convergent evolution and demonstrates the development of functional similarities derived from common constraints.

Corneal eyes
      When vertebrates emerged onto land, they acquired a new refracting surface, the cornea. Because of the difference in refractive index between air and water, a curved cornea is an image-forming lens in its own right. Its focal length is given by f = nr/(n-1), where n is the refractive index of the fluid of the eye, and r is the radius of curvature of the cornea. All land vertebrates have lenses, but the lens is flattened and weakened compared with a fish lens. In the human eye (eye, human) the cornea is responsible for about two-thirds of the eye's optical power, and the lens provides the remaining one-third.

      Spherical corneas, similar to spherical lenses, can suffer from spherical aberration. To avoid this, the human cornea developed an ellipsoidal shape, with the highest curvature in the centre. A consequence of this nonspherical design is that the cornea has only one axis (optical axis) of symmetry, and the best image quality occurs close to this axis, which corresponds with central vision (as opposed to peripheral vision). In addition, central vision is aided by a region of high photoreceptor density, known as the fovea or the less clearly defined “area centralis,” that lies close to the central axis of the eye and specializes in acute vision.

 Corneal eyes are found in spiders (spider), many of which have eyes with excellent image-forming capabilities. Spiders typically have eight eyes, two of which, the principal eyes, point forward and are used in tasks such as the recognition of members of their own species. Hunting spiders use the remaining three pairs, secondary eyes, as movement detectors. However, in web-building spiders, the secondary eyes are underfocused and are used as navigation aids, detecting the position of the Sun and the pattern of polarized light in the sky. Jumping spiders (jumping spider) have the best vision of any spider group, and their principal eyes can resolve a few minutes of arc, which is many times better than the eyes of the insects (insect) on which they prey. The eyes of jumping spiders are also unusual in that the retinas scan to and fro across the image while the spider identifies the nature of its target.

      Insects (insect) also have corneal single-chambered eyes. The main eyes of many insect larvae consist of a small number of ocelli, each with a single cornea. The main organs of sight of most insects as adults are the compound eyes, but flying insects also have three simple dorsal ocelli. These are generally underfocused, giving blurred images; their function is to monitor the zenith and the horizon, supplying a rapid reaction system for maintaining level flight.

Concave mirror eyes
      Scallops (scallop) (Pecten) have about 50–100 single-chambered eyes in which the image is formed not by a lens but by a concave mirror. In 1965 British neurobiologist Michael F. Land (the author of this article) found that although scallop eyes have a lens, it is too weak to produce an image in the eye. In order to form a visible image, the back of the eye contains a mirror that reflects light to the photoreceptors. The mirror in Pecten is a multilayer structure made of alternating layers of guanine and cytoplasm, and each layer is a quarter of a wavelength (about 0.1 μm in the visible spectrum) thick. The structure produces constructive interference for green light, which gives it its high reflectance. Many other mirrors in animals are constructed in a similar manner, including the scales of silvery fish, the wings of certain butterflies (butterfly) (e.g., the Morpho genus), and the iridescent (iridescence) feathers of many birds. The eyes of Pecten also have two retinas, one made up of a layer of conventional microvillus receptors close to the mirror and out of focus, and the second made up of a layer with ciliary receptors in the plane of the image. The second layer responds when the image of a dark object moves across it; this response causes the scallop to shut its shell in defense against potential predation.

      Reflecting eyes such as those of Pecten are not common. A number of copepod and ostracod (mussel shrimp) crustaceans possess eyes with mirrors, but the mirrors are so small that it is difficult to tell whether the images are used. An exception is the large ostracod Gigantocypris, a creature with two parabolic reflectors several millimetres across. It lives in the deep ocean and probably uses its eyes to detect bioluminescent organisms on which it preys. The images are poor, but the light-gathering power is enormous. A problem with all concave mirror eyes is that light passes through the retina once, unfocused, before it returns, focused, from the mirror. As a result, photoreceptors see a low-contrast image, and this design flaw probably accounts for the rare occurrence of these eyes.

Compound eyes
      Compound eyes are made up of many optical elements arranged around the outside of a convex supporting structure. They fall into two broad categories with fundamentally different optical mechanisms. In apposition compound eyes each lens with its associated photoreceptors is an independent unit (the ommatidium), which views the light from a small region of the outside world. In superposition eyes the optical elements do not act independently; instead, they act together to produce a single erect image lying deep in the eye. In this respect they have more in common with single-chambered eyes, even though the way the image is produced is quite different.

Apposition eyes
 Apposition eyes were almost certainly the original type of compound eye and are the oldest fossil eyes known, identified from the trilobites (trilobite) of the Cambrian Period, about 542–488 million years ago. Although compound eyes are most often associated with the arthropods (arthropod), especially insects (insect) and crustaceans (crustacean), compound eyes evolved independently in two other phyla, the mollusks (mollusk) and the annelids (annelid). In the mollusk phylum, clams (clam) of the genera Arca and Barbatia have numerous tiny compound eyes, each with up to a hundred ommatidia, situated around their mantles (mantle). In these tiny eyes each ommatidium consists of a photoreceptor cell and screening pigment cells. The eyes have no lenses (lens) and rely simply on shadowing from the pigment tube to restrict the field of view. In the annelid phylum the tube worms (tube worm) of the family Sabellidae have eyes similar to those of Arca and Barbatia at various locations on the tentacles. However, these eyes differ in that they have lenses. The function of the eyes of both mollusks and annelids (annelid) is much the same as the mirror eyes of Pecten; they see movement and initiate protective behaviour, causing the shell to shut or the organism to withdraw into a tube.

Image formation
      In arthropods most apposition eyes have a similar structure. Each ommatidium consists of a cornea, which in land insects is curved and acts as a lens. Beneath the cornea is a transparent crystalline cone through which rays converge to an image at the tip of a receptive structure, known as the rhabdom. The rhabdom is rodlike and consists of interdigitating fingerlike processes (microvilli) contributed by a small number of photoreceptor cells. The number of microvilli varies, with eight being the typical number found in insects (insect). In addition, there are pigment cells of various kinds that separate one ommatidium from the next; these cells may act to restrict the amount of light that each rhabdom receives. Beneath the photoreceptor cells there are usually three ganglionic layers—the lamina, the medulla, and the lobula—that form a set of neuronal relays, and the rhabdom is connected to these layers by a single axon. The neuronal relays map and remap input from the retinal photoreceptors, thereby generating increasingly complex responses to contrast, motion, and form.

      In aquatic insects and crustaceans the corneal surface cannot act as a lens because it has no refractive power. Some water bugs (e.g., Notonecta, or back swimmers (back swimmer)) use curved surfaces behind and within the lens to achieve the required ray bending, whereas others use a structure known as a lens cylinder. Similar to fish lenses, lens cylinders bend light, using an internal gradient of refractive index, highest on the axis and falling parabolically to the cylinder wall. In the 1890s Austrian physiologist Sigmund Exner was the first to show that lens cylinders can be used to form images in the eye. He discovered this during his studies of the ommatidia of the horseshoe crab Limulus.

      A problem that remained poorly understood until the 1960s is the relationship between the inverted images formed in individual ommatidia and the image formed across the eye as a whole. The question was first raised in the 1690s when Dutch scientist Antonie van Leeuwenhoek (Leeuwenhoek, Antonie van) observed multiple inverted images of his candle flame through the cleaned cornea of an insect eye. Later investigations of the ommatidial structure revealed that in apposition eyes each ommatidium is independent and sees a small portion of the field of view. The field of view is defined by the lens, which also serves to increase the amount of light reaching the rhabdom. Each rhabdom scrambles and averages the light it receives, and the individual ommatidial images are sent via neurons from the ommatidia to the brain. In the brain, the separate images are perceived as a single overall image. The array of images formed by the convex sampling surface of the apposition compound eye is functionally equivalent to the concave sampling surface of the retina in a single-chambered eye.

Neural superposition eyes
      Conventional apposition eyes, such as those of bees (bee) and crabs (crab), have a similar optical design to the eyes of flies (Diptera (dipteran)). However, in fly eyes the photopigment-bearing membrane regions of the photoreceptors are not fused into a single rhabdom. Instead, they stay separated as eight individual rodlets (effectively seven, since two lie one above the other), known as rhabdomeres, each with its own axon. This means that each ommatidium should be capable of a seven-point resolution of the image, which raises the problem of incorporating multiple inverted images into a single erect image that the ordinary apposition eye avoids. In 1967 German biologist Kuno Kirschfeld showed that the angles between the individual rhabdomeres in one ommatidium are the same as those between adjacent ommatidia. As a result, each of the seven rhabdomeres in one ommatidium shares a field of view with a rhabdomere in a neighbouring ommatidium. In addition, all seven rhabdomeres that share a common field of view send their axons (axon) to the same place in the first ganglionic layer—the lamina. Thus, at the level of the lamina the image is no different from that in an ordinary apposition eye. However, because each of the seven photoreceptor axon inputs connects to second-order neurons (neuron), the image at the level of the lamina is effectively seven times brighter than in the photoreceptors themselves. This allows flies to fly earlier in the morning and later in the evening than other insects with eyes of similar resolution. This variant of the apposition eye has been called neural superposition.

Wavelength and plane of polarization
      Although there is no further spatial resolution within a rhabdom, the various photoreceptors in each ommatidium do have the capacity to resolve two other features of the image, wavelength and plane of polarization. The different photoreceptors do not all have the same spectral sensitivities (sensitivities to different wavelengths). For example, in the honeybee there are three photopigments in each ommatidium, with maximum sensitivities in the ultraviolet, the blue, and the green regions of the spectrum. This forms the basis of a trichromatic colour vision system that allows bees to distinguish accurately between different flower colours. Some butterflies have four visual pigments (visual pigment), one of which is maximally sensitive to red wavelengths. The most impressive array of pigments is found in mantis shrimps (mantis shrimp) (order Stomatopoda), where there are 12 visual pigments in a special band across the eye. Eight pigments cover the visible spectrum, and four cover the ultraviolet region.

      Unlike humans, many arthropods (arthropod) have the ability to resolve the plane of polarized light. Single photons (photon) of light are wave packets in which the electrical and magnetic components of the wave are at right angles. The plane that contains the electrical component is known as the plane of polarization. sunlight contains photons polarized in all possible planes and therefore is unpolarized. However, the atmosphere scatters light selectively, in a way that results in a pattern of polarization in the sky that is directly related to the position of the Sun. Austrian zoologist Karl von Frisch (Frisch, Karl von) showed that bees could navigate by using the pattern of polarization instead of the Sun when the sky was overcast. The organization of the photopigment molecules on the microvilli in the rhabdoms (rhabdom) of bees makes this type of navigation possible. A photon will be detected only if the light-sensitive double bond of the photopigment molecule lies in the plane of polarization of the photon. The rhabdoms in the dorsal regions of bee eyes have their photopigment molecules aligned with the axes of the microvilli, which lie parallel to one another in the photoreceptor. As a result, each photoreceptor is able to act as a detector for a particular plane of polarization. The whole array of detectors in the bee's eyes is arranged in a way that matches the polarization pattern in the sky, thus enabling the bee to easily detect the symmetry plane of the pattern, which is the plane containing the Sun.

 The other physical process that results in polarization is reflection. For example, a water surface polarizes reflected light so that the plane of polarization is parallel to the plane of the surface. Many insects, including back swimmers of Notonecta, make use of this property to find water when flying between pools. The mechanism is essentially the same as in the bee eye. There are pairs of photoreceptors with opposing microvillar orientations in the downward-pointing region of the eye, and when the photoreceptors are differentially stimulated by the polarized light from a reflecting surface, the insect makes a dive. The reason that humans cannot detect polarized light is that the photopigment molecules can take up all possible orientations within the disks of the rods (rod) and cones (cone), unlike the microvilli of arthropods, in which the molecules are constrained to lie parallel to the microvillar axis.

Differences in resolution
      The number of ommatidia in apposition eyes varies from a handful, as in primitive wingless insects and some ants (ant), to as many as 30,000 in each eye of some dragonflies (dragonfly) (order Odonata). The housefly has 3,000 ommatidia per eye, and the vinegar fly (or fruit fly) has 700 per eye. In general, the resolution of the eye increases with increasing ommatidial number. However, the physical principle of diffraction means that the smaller the lens, the worse the resolution of the image. This is why astronomical telescopes have huge lenses (or mirrors), and it is also why the tiny lenses of compound eyes have poor resolution. A bee's eye, with 25-μm- (0.001-inch-) wide lenses, can resolve about one degree. The human eye, with normal visual acuity (20/20 vision), can resolve lines spaced less than one arc minute (one-sixtieth of one degree) apart, which is about 60 times better than a bee. In addition, the single lens of the human eye has an aperture (relative aperture) diameter (in daylight) of 2.5 mm (0.1 inch), 100 times wider than that of a single lens of a bee. If a bee were to attempt to improve its resolution by a factor of two, it would have to double the diameter of each lens, and it would need to double the number of ommatidia to exploit the improved resolution. As a result, the size of an apposition eye would increase as the square of the required resolution, leading to absurdly large eyes. In 1894 British physicist Henry Mallock calculated that a compound eye with the same resolution as human central vision would have a radius of 6 metres (19 feet). Given this problem, a resolution of one-quarter of a degree, found in the large eyes of dragonflies, is probably the best that any insect can manage.

      Because increased resolution comes at a very high cost in terms of overall eye size, many insects have eyes with local regions of increased resolution (acute zones), in which the lenses are larger. The need for higher resolution is usually connected with sex or predation. In many male dipteran flies and male (drone) bees, there is an area in the upper frontal region of the eyes where the facets are enlarged, giving resolution that is up to three times more acute than elsewhere in the eye. The acute resolution is used in the detection and pursuit of females. In one hover fly genus (Syritta) the males make use of their superior resolution to stay just outside the distance at which females can detect them. In this way a male can stalk a female on the wing until she lands on a flower, at which point he pounces. In a few flies, such as male bibionids (March flies (March fly)) and simuliids (black flies (black fly)), the high- and low-resolution parts of the eye form separate structures, making the eye appear doubled. Insects that catch other insects on the wing also have special “acute zones.” Both sexes of robber fly (family Asilidae) have enlarged facets in the frontal region of the eye, and dragonflies have a variety of more or less upward-pointing high-resolution regions that they use to spot flying insects against the sky. The hyperiid amphipods (amphipod), medium-sized crustaceans from the shallow and deep waters of the ocean, have visual problems similar to those of dragonflies, although in this case they are trying to spot the silhouettes of potential prey against the residual light from the surface. This has led to the development of highly specialized divided eyes in some species, most notably in Phronima, in which the whole of the top of the head is used to provide high resolution and sensitivity over a narrow (about 10 degrees) field of view. Not all acute zones are upward-pointing. Some empid flies (or dance flies (dance fly)), which cruise around just above ponds looking for insects trapped in the water surface, have enlarged facets arranged in a belt around the eye's equator—the region that views the water surface.

Superposition eyes
      Crepuscular (active at twilight) and nocturnal insects (e.g., moths (moth)), as well as many crustaceans (crustacean) from the dim midwater regions of the ocean, have compound eyes known as superposition eyes, which are fundamentally different from the apposition type. Superposition eyes look superficially similar to apposition eyes in that they have an array of facets around a convex structure. However, outside of this superficial resemblance, the two types differ greatly. The key anatomical features of superposition eyes include the existence of a wide transparent clear zone beneath the optical elements and a deep-lying retinal layer, usually situated about halfway between the eye surface and the centre of curvature of the eye. Unlike apposition eyes, where the lenses each form a small inverted image, the optical elements in superposition eyes form a single erect image, located deep in the eye on the surface of the retina. The image is formed by the superimposed (hence the name superposition) ray-contributions from a large number of facets. Thus, in some ways this type of eye resembles the single-chambered eye in that there is only one image, which is projected through a transparent region onto the retina.

Refracting, reflecting, and parabolic optical mechanisms
      In superposition eyes the number of facets that contribute to the production of a single image depends on the type of optical mechanism involved. There are three general mechanisms, based on lenses (refracting superposition), mirrors (reflecting superposition), and lens-mirror combinations (parabolic superposition).

      The refracting superposition mechanism was discovered by Austrian physiologist Sigmund Exner in the 1880s. He reasoned that the geometrical requirement for superposition was that each lens element should bend light in such a way that rays entering the element at a given angle to its axis would emerge at a similar angle on the same side of the axis. Exner realized that this was not the behaviour of a normal lens, which forms an image on the opposite side of the axis from the entering ray. He worked out that the only optical structures capable of producing the required ray paths were two-lens devices, specifically two-lens inverting telescopes. However, the lenslike elements of superposition eyes lack the necessary power in their outer and inner refracting surfaces to operate as telescopes. Exner solved this by postulating that the elements have a lens cylinder structure with a gradient of refractive index capable of bending light rays continuously within the structure. This is similar to the apposition lens cylinder elements in the Limulus eye (see above Apposition eyes (photoreception)); the difference is that the telescope lenses would be twice as long. The lens cylinder arrangement produces the equivalent of a pair of lenses, with the first lens producing a small image halfway down the structure and the second lens turning the image back into a parallel beam. In the process the ray direction is reversed. Thus, the emerging beam is on the same side of the axis as the entering beam—the condition for obtaining a superposition image from the whole array. In the 1970s, studies using an interference microscope, a device capable of exploring the refractive index distribution in sections of minute objects, showed that Exner's brilliant idea was accurate in all important details.

      There is one group of animals with eyes that fit the anatomical criteria for superposition but that have optical elements that are not lenses or lens cylinders. These are the long-bodied decapod crustaceans, such as shrimps (shrimp), prawns (prawn), crayfish, and lobsters (lobster). The optical structures are peculiar in that they have a square rather than a circular cross section, and they are made of homogeneous low-refractive index jelly. For a period of 20 years—between 1955, when interference microscopy showed that the jelly structures lacked appropriate refracting properties, and 1975, when the true nature of these structures was discovered—there was much confusion about how these eyes might function. Working with crayfish eyes, German neurobiologist Klaus Vogt found that these unpromising jelly boxes were silvered with a multilayer reflector coating. A set of plane mirrors, aligned at right angles to the eye surface, change the direction of rays (in much the same way as len cylinders), thereby producing a single erect image by superposition. The square arrangement of the mirrors has particular significance. Rays entering the eye at an oblique angle encounter two surfaces of each mirror box rather than one surface. In this case, the pair of mirrors at right angles acts as a corner reflector. Corner reflectors reflect an incoming ray through 180 degrees, irrespective of the ray's original direction. As a result, the reflectors behave as though they were a single plane mirror at right angles to the ray. This ensures that all parallel rays reach the same focal point and means that the eye as a whole has no single axis, which allows the eye to operate over a wide angle.

      The third type of superposition eye, discovered in 1988 in the crab genus Macropipus by Swedish zoologist Dan-Eric Nilsson, has optical elements that use a combination of a single lens and a parabolic mirror. The lens focuses an image near the top of the clear zone (similar to an apposition eye), but oblique rays are intercepted by a parabolic mirror surface that lines the crystalline cone beneath the lens. The parabolic mirror unfocuses the light and redirects it back across the axis of the structure, producing an emerging ray path similar to that of a refracting or reflecting superposition eye.

      All three types of superposition eyes have adaptation mechanisms that restrict the amount of light reaching the retina in bright conditions. In most cases, light is restricted by the migration of dark pigment (held between the crystalline cones in the dark) into the clear zone; this cuts off the most oblique rays. However, as the pigment progresses inward, it cuts off more and more of the image-forming beam until only the central optical element supplies light to the rhabdom (located immediately below the central optical element). This effectively converts the superposition eye into an apposition eye, since in the dark-adapted condition up to a thousand facets may contribute to the image at any one point on the retina, potentially reducing the retinal illumination a thousandfold.

optics of superposition eyes
      Superposition optics requires that parallel rays from a large portion of the eye surface meet at a single point in the image. As a result, superposition eyes should have a simple spherical geometry, and, in fact, most superposition eyes in both insects and crustaceans are spherical. Some moth eyes do depart slightly from a spherical form, but it is in the euphausiid crustaceans ( krill) from the mid-waters of the ocean that striking asymmetries are found. In many krill species the eyes are double. One part, with a small field of view, points upward, and a second part, with a wide field of view, points downward (similar to the apposition eyes of hyperiid amphipods). It is likely that the upper part is used to spot potential prey against the residual light from the sky, and the lower part scans the abyss (abyssal zone) for bioluminescent organisms. The most extraordinary double superposition eyes occur in the tropical mysid shrimp genus Dioptromysis, which has a normal-looking eye that contains a single enormous facet embedded in the back, with an equally large lens cylinder behind the facet. This single optical element supplies a fine-grain retina, which seems to act as the “fovea” of the eye as a whole. At certain times the eyes rotate so that the single facets are directed forward to view the scene ahead with higher resolution, much as one would use a pair of binoculars (binocular).

Structure and function of photoreceptors
      Photoreceptors are the cells (cell) in the retina that respond to light. Their distinguishing feature is the presence of large amounts of tightly packed membrane that contains the photopigment rhodopsin or a related molecule. The tight packing is needed to achieve a high photopigment density, which allows a large proportion of the light photons (photon) that reach the photoreceptor to be absorbed. Photon absorption contributes to the photoreceptor's output signal.

      In the retina of vertebrates (vertebrate) the rods (rod) and cones (cone) have photopigment-bearing regions (outer segments) composed of a large number of pancakelike disks. In rods the disks are closed, but in cones the disks are partially open to the surrounding fluid. In a typical rod there are about a thousand disks, and each disk holds about 150,000 rhodopsin molecules, giving a total of 150 million molecules per rod. In most invertebrate photoreceptors the structure is different, with the photopigment borne on regularly arranged microvilli, fingerlike projections with a diameter of about 0.1 μm. This photoreceptor structure is known as a rhabdom. The photopigment packing is less dense in rhabdoms than in vertebrate disks. In both vertebrate photoreceptors and rhabdoms, each photoreceptor cell contains a nucleus, an energy-producing region with mitochondria (mitochondrion) (in the inner segment in rods and cones), and an axon that conveys electrical signals to the next neurons (neuron) in the processing chain. In reptiles (reptile) and birds (bird) the receptors may also contain coloured oil droplets that modify the spectrum of the light absorbed by the photopigment, thereby enhancing colour vision. In insects and other invertebrates the receptors may also contain granules of dark pigment that move toward the rhabdom in response to light. They act as a type of pupil, protecting the rhabdom in bright conditions by absorbing light.

Photopigments
      The photopigments that absorb light (photochemical reaction) all have a similar structure, which consists of a protein called an opsin and a small attached molecule known as the chromophore. The chromophore absorbs photons of light, using a mechanism that involves a change in its configuration. In vertebrate rods the chromophore is retinal, the aldehyde of 1 (vitamin A). When retinal absorbs a photon, the double bond between the 11th and 12th carbon atoms flips, thus reconfiguring the molecule from the 11-cis to the all-trans form. This in turn triggers a molecular transduction cascade, resulting in the closure of sodium channels in the membrane and hyperpolarization (increase in negativity) of the cell. Retinal then detaches from opsin, is regenerated to the 11-cis state in the cells of the pigment epithelium that surround the rods, and is reattached to an opsin molecule. In most invertebrate photoreceptors the chromophore does not detach from opsin but is regenerated in situ, usually by the absorption of a photon with a wavelength different from the stimulating wavelength.

      The opsin molecules themselves each consist of seven helices that cross the disk membrane and surround the chromophore. Humans have four different opsins. One type is found in rods and is responsible for low-light vision, and three types are found in cones and subserve colour vision by responding to blue, green, and red wavelengths. The differences in the amino acid compositions of the opsins have the effect of altering the charge environment around the chromophore group, which in turn shifts the wavelength to the photopigment that is maximally sensitive. Thus, in humans the rods are most sensitive to light in the blue-green spectrum (peak wavelength 496 nm), and the cones are most sensitive to light in the blue (419 nm), green (531 nm), and yellow-green (or red; 558 nm) spectra. The cones are often designated as short (S), medium (M), and long (L) wavelength cones.

      Most perceived colours are interpreted by the brain from a ratio of excitation in different cone types. The fact that the spectral sensitivity maxima of the M and L cones are very close together reveals an interesting evolutionary history. Most fish and birds have four or even five cone types with different spectral sensitivities, including sensitivity in the ultraviolet (ultraviolet radiation). In contrast, most mammals (mammal) have only two—an S cone for blue wavelengths and an L cone for red wavelengths. Thus, these mammals have dichromatic vision, and they are red-green colour-blind. The relative poverty of the mammalian colour system is probably due to the way that the early mammals survived the age of reptiles by adopting a nocturnal and even subterranean way of life in which colour vision was impossible. However, about 63 million years ago a mutation in the genotype of the Old World primates (primate) resulted in the duplication of the gene for the long-wavelength opsin, which provided another channel for a trichromatic colour vision system. The red-green system of M and L cones enabled primates to distinguish particular elements in their environment—for example, the ripeness of fruit in the tropical woodlands that the early primates inhabited.

      Retinal is not the only chromophore of rhodopsins (rhodopsin); for example, vertebrates have another chromophore, 3-dehydroretinal, which gives rise to a family of photopigments known as porphyropsins. Relative to retinal-based pigments with the same opsin, the spectral sensitivity of porphyropsins is shifted about 30 nm toward the red end of the spectrum. Other chromophores include 3-hydroxyretinal, which is present in some insects and produces a photopigment known as xanthopsin, and 4-hydroxyretinal, which is present in the firefly squid (Watasenia). Firefly squid appear to have a colour vision system that is based on photopigments with the same opsin but with three different chromophores. In most other colour vision systems (including all the visual pigments (visual pigment) in humans), the chromophore stays the same, and spectral tuning is achieved by varying the amino acid composition of the opsins.

Neural transmission
 All vertebrates have complex retinas (retina) with five layers, first described in detail by Spanish histologist Santiago Ramón y Cajal (Ramón y Cajal, Santiago) in the 1890s. There are three layers of cells (cell) on the pathway from the photoreceptors to the optic nerve. These are the photoreceptors themselves at the rear of the retina, the bipolar cells, and finally the ganglion cells, whose axons (axon) make up the optic nerve. Forming a network between the photoreceptors and the bipolar cells are the horizontal cells (the outer plexiform layer), and between the bipolar cells and the ganglion cells, there exists a similar layer (the inner plexiform layer) containing amacrine cells of many different kinds. A great deal of complex processing occurs within the two plexiform layers. The main function of the horizontal cells is to vary the extent of coupling between photoreceptors and between photoreceptors and bipolar cells. This provides a control system that keeps the activity of the bipolar cells within limits, regardless of fluctuations in the intensity of light reaching the receptors. This control process also enhances contrast, thus emphasizing the differences between photoreceptor outputs.

      The bipolar cells are of two kinds—“on” and “off”—responding to either an increase or a decrease in local light intensity. The roles of the amacrine cells are less clear, but they contribute to the organization of the receptive fields of the ganglion cells. These fields are the areas of retina over which the cells respond. Typically, receptive fields have a concentric structure made up of a central region surrounded by an annular ring, with the central and annular areas having opposite properties. Thus, some ganglion cells are of the “on-centre/off-surround” type, and others are of the “off-centre/on-surround” type. In practical terms, this means that a small contrasting object crossing the receptive field centre will stimulate the cell strongly, but a larger object, or an overall change in light intensity, will not stimulate the cell, because the effects of the centre region and annular ring cancel one another. Thus, ganglion cells are detectors of local contrast rather than light intensity. Many ganglion cells in primates (primate) also show colour opponency—for example, responding to “red-on/green-off” or “blue-on/yellow-off” and signaling information about the wavelength structure of the image. Thus, in the stages of processing an image, the components of contrast, change, and movement appear to be the most biologically important.

      In the vertebrate retina a series of biochemical stages (chemical reaction) convert the isomerization of the retinal of the rhodopsin molecule (from 11-cis to all trans) into an electrical signal. Within about one millisecond of photon absorption, the altered rhodopsin molecule becomes excited, causing activation of a heterotrimeric G-protein ( guanine nucleotide binding protein) called transducin. G-proteins act as mediators of cell signaling pathways that involve relay signaling molecules called second messengers (second messenger). In the case of rhodopsin excitation, transducin activates an enzyme called phosphodiesterase, which cleaves a second messenger known as cGMP (3′5′-cyclic guanosine monophosphate) into 5′GMP. This process reduces the amount of cGMP in the cell.

      In dark conditions, cGMP binds to sodium channels in the cell membrane (cell), keeping the channels open and allowing sodium ions (ion) to enter the cell continuously. The constant influx of positive sodium ions maintains the cell in a somewhat depolarized (weakly negative) state. In light conditions, cGMP does not bind to the channels, which allows some sodium channels to close and cuts off the inward flow of sodium ions. The reduction in influx of sodium ions causes the cell to become hyperpolarized (strongly negative). Thus, the electrical effect of a photon of light is to cause a short-lived negative potential in the photoreceptor. Bright light produces more rhodopsin isomerizations, further decreasing cGMP levels and enabling hyperpolarization to be graded with light intensity. The electrical signal produced by light reaches the base of the inner segment of the receptor, where a neuronal synapse releases vesicles of neurotransmitter (in this case glutamate) in proportion to voltage in the receptor.

      In invertebrate eyes the electrical response to light is different. The majority of invertebrate eyes have microvillus receptors that depolarize (become less negative) when illuminated—the opposite of the response in vertebrate receptors. The depolarization is brought about by the entry of sodium and calcium ions that results from the opening of membrane channels. The biochemistry of the transducer pathway is not entirely clear; some proposed models envision a somewhat different pathway from that in vertebrates. Rhodopsin isomerization activates a G-protein, which in turn activates an enzyme called phospholipase C (PLC). PLC catalyzes the production of an intracellular second messenger known as IP3 (inositol 1,4,5-trisphosphate), which stimulates the release of calcium from intracellular stores in certain organelles. It is not entirely clear what causes the membrane channels to open; however, there is evidence that calcium plays a major role in this process. In contrast to other invertebrates, the “off”-responding distal receptors of the scallop retina work by a different mechanism. They hyperpolarize to light (similar to vertebrate receptors) by closing sodium channels, which also results in the simultaneous release of potassium ions from cells.

Adaptive mechanisms of vision
      The human visual system manages to provide a usable signal over a broad range of light intensities. However, some eyes (eye, human) are better adapted optically to dealing with light or dark conditions. For example, the superposition eyes of nocturnal moths (moth) may be as much as a thousand times more sensitive than the apposition eyes of diurnal butterflies (butterfly). Within vertebrate eyes, there are four kinds of mechanisms that operate to allow vision across a wide range of light intensities. These include mechanisms specific to the iris, the splitting of the intensity range between rods (rod) and cones (cone), adjustments to the signal transduction process in the photoreceptors, and variations in the availability of active photopigment molecules.

Vision and light intensity
      The most obvious mechanism involved in light regulation is the iris. In humans the iris opens in the dark to a maximum diameter of 8 mm (0.31 inch) and closes to a minimum of 2 mm (0.08 inch). The image brightness in the retina changes by a factor of 16. In other animals the effect of the pupil may be much greater; for example, in certain geckos (gecko) the slit pupil can close from a circle of several millimetres in diameter down to four pinholes each, with a diameter of 0.1 mm (0.004 inch) or less. The retinal brightness ratio is at least a thousandfold. The reason for this great range is probably that the gecko's nocturnal eye needs strong protection from bright daylight.

      In humans the rods are concerned with the dimmest part of the eye's working range and have no colour vision. The cones begin to take over at about the level of bright moonlight, and at all daylight intensities the cones alone provide the visual signal. Rods respond to single photons (photon) of light with large electrical signals, which means that the electrical responses saturate at low rates of photon capture by the rhodopsin molecules. Rods operate over the range from the threshold of vision, when they are receiving about one photon every 85 minutes, to dawn and dusk conditions, when they receive about 100 photons per second. For most of their range the rods are signaling single photon captures. The cones are much less sensitive than the rods; they still respond to single photons, but the sizes of the resulting electrical signals are much smaller. This gives the cones a much larger working range, from a minimum of about three photons per second to more than a million per second, which is enough to deal with the brightest conditions that humans encounter.

      If cones are presented with brief flashes, rather than steady illumination changes, their working range from threshold to saturation is small—reduced to a factor of about 100. However, longer illumination induces two kinds of change that extend this range. The biochemical transducer cascade that leads to the electrical signal has an ability to regulate its own gain, thereby reducing the size of the electrical signal at high photon capture rates. The main mechanism depends on the fact that calcium ions, which enter the photoreceptor along with sodium ions, have an inhibitory effect on the synthesis of cGMP, the molecule that keeps the sodium channels open (see above Structure and function of photoreceptors: Neural transmission (photoreception)). The effect of light is to reduce cGMP levels and thus close the membrane channels to sodium and calcium. If the light is persistent, calcium levels in the photoreceptor fall, the calcium “brake” on cGMP production weakens, and cGMP levels increase somewhat. Increased cGMP production opens the membrane channels again. Thus, there is a feedback loop that tends to oppose the direct effect of light, ensuring that saturation (complete closure of all the membrane channels) does not occur. This in turn extends the top end of the photoreceptor's working range.

      The slow speed of turnover of functional visual pigment molecules also helps to extend the eye's ability to respond to high light levels. In vertebrates the all-trans retinal, produced when a photon isomerizes the 11-cis retinal of a rhodopsin molecule, is removed from the rod or cone. It passes to the adjacent pigment epithelium, where it is regenerated back to the active 11-cis form and passed back to the photoreceptor. On average, this process takes two minutes. The higher the light level, the greater the number of molecules of retinal in the inactive all-trans state. Therefore, there are fewer rhodopsin molecules available to respond to light. At the top end of the intensity distribution, photoreception becomes self-limiting, with the cones never catching more than about one million photons per second.

Eye movements and active vision
 There are four main types of eye movement: saccades (saccade), reflex stabilizing movements, pursuit movements, and vergence movements. Saccades are fast movements that redirect gaze. They may involve the eyes alone or, more commonly, the eyes and the head. Their function is to place the fovea (the central region of the retina where vision is most acute) onto the images of parts of the visual scene of interest. The duration and peak velocity of saccades vary systematically with their size. The smallest movements, microsaccades, move the eye through only a few minutes of arc. They last about 20 milliseconds and have maximum velocities of about 10 degrees per second. The largest saccades (excluding the contributions of head movements) can be up to 100 degrees, with a duration of up to 300 milliseconds and a maximum velocity of about 500–700 degrees per second. During saccades, vision is seriously impaired for two reasons. First, during large saccades, the image is moving so fast that it is blurred and unusable. Second, an active blanking-off process, known as saccadic suppression, occurs, and this blocks vision for the first part of each saccade. Between saccades, the eyes are held stationary in fixations. It is during these periods, which last on average about 190 milliseconds, that the eye takes in visual information. Saccades can be reflexive in nature—for example, when an object appears in one's peripheral field of view. However, as Russian psychologist Alfred L. Yarbus showed, saccades are often information-seeking in nature, directed to particular objects or regions by the requirements of ongoing behaviour.

      During fixations the eyes are stabilized against movements of the head and body by two reflexes, the vestibulo-ocular reflex (VOR) and the optokinetic reflex (OKR). In VOR the semicircular canals of the inner ear measure rotation of the head and provide a signal for the oculomotor nuclei of the brainstem, which innervate the eye muscles (muscle). The muscles counterrotate the eyes in such a way that a rightward head rotation causes an equal leftward rotation of both eyes, with the result that gaze direction stays stationary. OKR is a feedback loop in which velocity-sensitive ganglion cells in the retina feed a signal, via the oculomotor nuclei, to the eye muscles. The effect of the feedback loop is to move the eye in the same direction as the image motion. With a moving background (e.g., when looking out of a train window), OKR ensures that the eye moves at almost the same speed as the image, and the result is optokinetic nystagmus, a sawtooth motion in which OKR alternates with saccadelike movements that reset the eyes to a central position. However, the principal function of OKR is to keep gaze stationary by nulling out any involuntary motion that results from visual drift or slow head movement. In general OKR and VOR work together to keep the image stationary on the retina, with VOR compensating for fast movements and OKR for slower movements.

      Humans and other primates (primate) have the ability to track moving objects with their eyes; this capacity is not widespread in mammals (mammal) or other vertebrates. These tracking movements employ a velocity feedback loop (similar to OKR) that functions only for small centrally placed targets (unlike OKR, which works over a much wider field). Smooth tracking, in which the eye moves continuously with the target, is typically confined to slow speeds (less than 20 degrees per second), although it sometimes can match targets moving up to 90 degrees per second. For faster objects the eye lags behind the target and catches up to it by using saccades. Thus, when watching a tennis match, the eyes track the ball with a mixed strategy of smooth movements and saccades.

      Vergence movements occur as an object approaches or recedes from the observer. They differ from other eye movements in that the two eyes move in opposite directions. Vergence movements are confined to humans and other animals with frontal eyes that employ binocular mechanisms to determine distance.

      The saccade-and-fixate strategy is the way humans take in information from the world most of the time. However, there is a mismatch between the extremely jerky movements of the image on the retina and the apparently smooth and coherent view of the world that is perceived consciously. While there is no scientific explanation for this discrepancy, it is clear that humans retain little information from one fixation to the next. If observers are presented with alternating views of the same scene, but with one substantial change between views, it takes many presentations before the change is detected if a blank period equivalent to a saccade is introduced between each view. However, if there is no blank period, the change is readily detected because it produces a visible local change in the image, which attracts attention. This phenomenon, known as change blindness, seems to imply that one reason humans do not “see” saccades is that the preceding image is not retained. Thus, humans have no basis for detecting the change that each saccade causes.

      At first sight the function of saccades and fixations appears to be to move the fovea from one interesting point in the scene to another. However, that is not how the saccades-fixation eye movement pattern originated. goldfish, which have no foveae, show the same saccades-fixation pattern as crabs (crab) and even cuttlefish, both of which have foveae. Flying houseflies (housefly) make head saccades (they do not have independently movable eyes) separated by stabilized periods. As American optometrist and physiologist Gordon Lynn Walls pointed out, the real significance of the saccades-fixation eye movement pattern is to keep gaze stationary. Saccades, on that basis, are simply a way of shifting the scene as fast as possible in order for vision to be lost for as short an amount of time as is practicable.

      Image movement also causes blur (i.e., loss of contrast in the finest detail of the image). Photoreception is a slow process, and it may take 20 milliseconds or more for a full response to a local change in light intensity to occur. This causes vision to become compromised. In humans the field of view of photoreceptors is 1 minute of arc; if an image moves faster on the retina than 1 minute in 20 milliseconds (0.83 degree per second), the finest detail in the image will begin to blur. This is a very slow speed and emphasizes the need for effective stabilizing mechanisms, such as VOR and OKR.

Central processing of visual information
 Vivid images of the world, with detail, colour, and meaning, impinge on human consciousness. Many people believe that humans simply see what is around them. However, internal images are the product of an extraordinary amount of processing, involving roughly half the cortex (the convoluted outer layer) of the brain. This processing does not follow a simple unitary pathway. It is known both from electrical recordings and from the study of patients with localized brain damage that different parts of the cerebral cortex abstract different features of the image; colour, depth, motion, and object identity all have “modules” of cortex devoted to them. What is less clear is how multiple processing modules assemble this information into a single image. It may be that there is no resynthesis, and what humans “see” is simply the product of the working of the whole visual brain.

      The axons (axon) of the ganglion cells leave the retina in the two optic nerves (optic nerve), which extend to the two lateral geniculate nuclei (LGN) in the thalamus. The LGN act as way stations on the pathway to the primary visual cortex, in the occipital (rear) area of the cerebral cortex. Some axons also go to the superior colliculus, a paired formation on the roof of the midbrain. Between the eyes and the lateral geniculate nuclei, the two optic nerves split and reunite in the optic chiasm, where axons from the left half of the field of view of both eyes join. From the chiasm the axons from the left halves make their way to the right LGN, and the axons from the right halves make their way to the left LGN. The significance of this crossing-over is that the two images of the same part of the scene, viewed by the left and right eyes, are brought together. The images are then compared in the cortex, where differences between them can be reinterpreted as depth in the scene. In addition, the optic nerve fibres have small, generally circular receptive fields with a concentric “on”-centre/“off”-surround or “off”-centre/“on”-surround structure. This organization allows them to detect local contrast in the image. The cells (cell) of the LGN, to which the optic nerve axons connect via synapses (synapse) (electrical junctions between neurons), have a similar concentric receptive field structure. A feature of the LGN that seems puzzling is that only about 20–25 percent of the axons reaching them come from the retina. The remaining 75–80 percent descend from the cortex or come from other parts of the brain. Some scientists suspect that the function of these feedback pathways may be to direct attention to particular objects in the visual field, but this has not been proved.

      The LGN in humans contain six layers of cells. Two of these layers contain large cells (the magnocellular [M] layers), and the remaining four layers contain small cells (the parvocellular [P] layers). This division reflects a difference in the types of ganglion cells that supply the M and P layers. The M layers receive their input from so-called Y-cells, which have fast responses, relatively poor resolution, and weak or absent responses to colour. The P layers receive input from X-cells, which have slow responses but provide fine-grain resolution and have strong colour responses. The division into an M pathway, concerned principally with guiding action, and a P pathway, concerned with the identities of objects, is believed to be preserved through the various stages of cortical processing.

      The LGN send their axons exclusively to the primary visual area (V1) in the occipital lobe of the cortex. The V1 contains six layers, each of which has a distinct function. Axons from the LGN terminate primarily in layers four and six. In addition, cells from V1 layer four feed other layers of the visual cortex. American biologist David Hunter Hubel (Hubel, David Hunter) and Swedish biologist Torsten Nils Wiesel (Wiesel, Torsten Nils) discovered in pioneering experiments beginning in the late 1950s that a number of major transformations occur as cells from one layer feed into other layers. Most V1 neurons (neuron) respond best to short lines and edges running in a particular direction in the visual field. This is different from the concentric arrangement of the LGN receptive fields and comes about by the selection of LGN inputs with similar properties that lie along lines in the image. For example, V1 cells with LGN inputs of the “on”-centre/“off”-surround type respond best to a bright central stripe with a dark surround. Other combinations of input from the LGN cells produce different variations of line and edge configuration. Cells with the same preferred orientation are grouped in columns that extend through the depth of the cortex. The columns are grouped around a central point, similar to the spokes of a wheel, and preferred orientation changes systematically around each hub. Within a column the responses of the cells vary in complexity. For example, simple cells respond to an appropriately oriented edge or line at a specific location, whereas complex cells prefer a moving edge but are relatively insensitive to the exact position of the edge in their larger receptive fields.

      Each circular set of orientation columns represents a point in the image, and these points are laid out across the cortex in a map that corresponds to the layout in the retina (retinotopic mapping). However, the cortical map is distorted compared with the retina, with a disproportionately large area devoted to the fovea and its immediate vicinity. There are two retinotopic mappings—one for each eye. This is because the two eyes are represented separately across the cortex in a series of “ocular dominance columns,” whose surfaces appear as curving stripes across the cortex. In addition, colour is carried not by the orientation column system but by a system prosaically known as “blobs.” These are small circular patches in the centre of each set of orientation columns, and their cells respond to differences in colour within their receptive fields; they do not respond to lines or edges.

      The processing that occurs in area V1 enables further analysis of different aspects of the image. There are at least 20 areas of cortex that receive input directly or indirectly from V1, and each of these has a retinotopic mapping. In front of V1 is V2, which contains large numbers of cells sensitive to the same features in each eye. However, within V2 there are small horizontal differences in direction between the eyes. Separation of the images in the two eyes results from the presence of objects in different depth planes, and it can be assumed that V2 provides a major input to the third dimension in the perceived world. Two other visual areas that have received attention are V4 and MT (middle temporal area, or V5). British neurobiologist Semir Zeki showed that V4 has a high proportion of cells that respond to colour in a manner that is independent of the type of illumination (colour constancy). This is in contrast to the cells of V1, which are responsive to the actual wavelengths (wavelength) present. In rare instances when V4 is damaged, the affected individual develops central achromatopsia, the inability to see or even imagine colours despite a normal trichromatic retina. Thus, it appears that V4 is where perceived colour originates. MT has been called the motion area, and cells respond in a variety of ways not only to movements of objects but also to the motion of whole areas of the visual field. When this area is damaged, the afflicted person can no longer distinguish between moving and stationary objects; the world is viewed as a series of “stills,” and the coordination of everyday activities that involve motion becomes difficult.

      In the 1980s American cognitive scientists Leslie G. Ungerleider and Mortimer Mishkin formulated the idea that there are two processing streams emanating from V1—a dorsal stream leading to the visual cortex of the parietal lobe and a ventral stream leading to the visual regions of the temporal lobe. The dorsal stream provides the parietal lobe with the position and information needed for the formulation of action; MT is an important part of this stream. The ventral stream is more concerned with detail, colour, and form and involves information from V4 and other areas. In the temporal lobe there are neurons with a wide variety of spatial form requirements, but these generally do not correspond exactly with any particular object. However, in a specific region of the anterior part of the inferotemporal cortex (near the end of the ventral stream) are neurons that respond to faces and very little else. Damage to areas near this part of the cortex can lead to prosopagnosia, the inability to recognize by sight people who are known to the subject. Loss of visual recognition suggests that information supplied via the ventral stream to the temporal lobe is refined and classified to the point where structures as complex as faces can be represented and recalled.

      Great progress has been made over the last century in understanding the ways that the eye and brain transduce and analyze the visual world. However, little is known about the relationship between the objective features of an image and an individual's subjective interpretation of the image. Scientists suspect that subjective experience is a product of the processing that occurs in the various brain modules contributing to the analysis of the image.

Evolution of eyes (eye, human)
 The soft-bodied animals (animal) that inhabited the world's seas before the Cambrian Period explosion (about 540 mya [million years ago]) undoubtedly had eyes (eye, human), probably similar to the pigment-pit eyes of flatworms (flatworm) today. However, there is no fossil evidence to support the presence of eyes in the early soft-bodied creatures. Scientists know that the photopigment rhodopsin existed in the Cambrian Period. Evidence for this comes from modern metazoan phyla, which have genetically related rhodopsins, even though the groups themselves diverged from a common ancestor well before the Cambrian.

 By the end of the early Cambrian Period (roughly 520 mya), most, if not all, of the eye types in existence today had already evolved. The need for better eyesight arose because some of the animals in the early Cambrian Fauna had turned from grazing to predation. Both predators and prey needed eyes to detect one another. Besides becoming better equipped visually, Cambrian animals developed faster forms of locomotion, and many acquired armoured exoskeletons (exoskeleton), which have provided fossil material. Many of the animals in the famous Burgess Shale deposits in British Columbia, Can., had convex eyes that presumably had a compound structure. The best-preserved compound eyes from the Cambrian Period are found in the trilobites (trilobite). Trilobite lenses were made of the mineral calcite, which enabled these organisms to fossilize exceptionally well. It is less certain when eyes of the camera-like single-chambered type first evolved. Fossil cephalopod mollusks (mollusk) appeared in the late Cambrian, and they probably had eyes resembling those of their present-day counterparts, such as lens eyes of octopus or pinhole eyes of nautilus.

      The first fish arose in the Ordovician Period (488.3–443.7 mya) and radiated extensively in the Devonian Period (416–359.2 mya). Fish fossils from these periods have eye sockets, indicating that these fish must have had eyes. The lampreys (lamprey), present-day relatives of these early fish, have eyes that are very similar to those of other fish, leading to the conclusion that very little has happened to the aquatic form of the vertebrate eye for about 400 million years. The lower chordates (chordate), from which the vertebrates arose, have either simple eyespots (eyespot) or no eyes at all; therefore, presumably the vertebrate eye originated with the first fish and not before.

      Given the short time that eyes had to evolve in the Cambrian Period (some estimates of the explosive phase of the Cambrian radiation are as short as 20 million years), it is of some interest to know how long it would actually take an eye to evolve. British naturalist Charles Darwin (Darwin, Charles) was concerned about the difficulty of evolving an eye because it was “an organ of extreme perfection and complication.” Thus, it might be expected that eye evolution would take a long time. In 1994 Swedish zoologists Dan-Eric Nilsson and Susanne Pelger took up the challenge of “evolving” an eye of the fish type from a patch of photosensitive skin. Using pessimistic estimates of variation, heritability, and selection intensity, Nilsson and Pelger came to the conclusion that it would take 364,000 generations for a fish eye to evolve. Given a generation time of a year, which is typical for moderate-sized animals, a respectable eye could evolve in less than half a million years. Of course, other physiological elements (e.g., competent brains (brain)) have to evolve in parallel with eyes. However, at least as far as the eye itself is concerned, very little time is actually required for its evolution.

      Another problem concerning the evolution of eyes is the number of times eyes evolved. Given that the fossil record does not contain much information about the eyes of Precambrian (Precambrian time) animals, scientists have had to rely on evidence from the eyes of living Precambrian descendants to solve this problem. In 1977 Austrian zoologist Luitfried von Salvini-Plawen and American biologist Ernst Mayr (Mayr, Ernst) examined the eyes and eyespots of representatives of all the main animal phyla and concluded that eyes of a basic kind had arisen independently at least 40 times and possibly as many as 65 times. The evidence presented by Salvini-Plawen and Mayr was of several kinds. At a cellular level, the receptive membrane of the photoreceptors could be elaborated from cilia (cilium) or from microvilli (fingerlike projections), the eyes could be derived either from epithelium or from nervous tissue, the axons (axon) of the receptors could leave from the back of the eye (everse) or from the front of the eye (inverse), and the overall eye design might be of the compound or the single-chambered type. Because these eye features tend to be stable within each phylum, the different combinations of features among phyla were taken to mean that the eyes had evolved independently. Set against this conclusion is the fact that some of the molecules involved in eye construction are indeed similar across phyla. The rhodopsin molecule itself is sufficiently similar among the vertebrates, the arthropods (arthropod), and the cephalopod mollusks (mollusk) to rely on common ancestry as the most likely explanation for similarity in eye construction. A gene that is associated with eye development, Pax-6 (paired box gene 6), is very similar in insects (insect) and mammals (mammal), and it also occurs in the eyes of cephalopod mollusks. Thus, the earliest metazoans had at least some of the molecules necessary for producing eyes. These molecules were passed on to the metazoans' descendants, who used the molecules in different ways to produce eyes of very varying morphology.

 Because there are only a limited number of ways that images can be produced, it is not surprising that some of them have been “discovered” more than once. This has led to numerous examples of convergence in the evolutionary history of eyes. The similarity in optical design of the eyes of fish and cephalopod mollusks, such as octopuses (octopus) and squid, is perhaps the most well-known example, but it is only one of many. The same lens design is also found in several groups of gastropod mollusks, in certain predatory worms (worm) (family Alciopidae), and in copepod crustaceans (genus Labidocera). A similar lens structure is also found in the extraordinary intracellular eye of a dinoflagellate protozoan (genus Warnowia). Compound eyes probably evolved independently in the chelicerata (genus Limulus), the trilobites, and the myriapods (myriapod) (genus Scutigera). Compound eyes appear to have evolved once or several times in the crustaceans and insects, in the bivalve mollusks (genus Arca), and in the annelid worms (genus Branchiomma). There are comparatively few cases in which one type of eye has evolved into a different type. However, it is thought that the single-chambered eyes of spiders (spider) and scorpions (scorpion) are descended from the compound eyes of earlier chelicerates (e.g., genus Limulus, eurypterids (eurypterid)) by a process of reduction. Something similar has occurred in the amphipod crustacean genus Ampelisca, where single-chambered eyes have replaced the compound eyes typical of the group.

Michael Land

Additional Reading

Photoreception and optical systems of eyes
A monumental ongoing series covering all aspects of sensory reception in organisms is Hansjochem Autrum (ed.), Handbook of Sensory Physiology (1971–). An introductory work pertaining specifically to photoreception is Robert W. Rodieck, The First Steps in Seeing (1998). The optical systems of eyes are discussed in relation to their role in vision in a wide range of organisms in Jerome J. Wolken, Light Detectors, Photoreceptors, and Imaging Systems in Nature (1995); and Michael F. Land and Dan-Eric Nilsson, Animal Eyes (2002). The types and functions of eye movements are covered in Roger H.S. Carpenter, Movements of the Eyes, 2nd ed. (1988). An appealing work on the basic aspects of the different eye structures and the mechanisms of photoreception specific to invertebrates is Eric Warrant and Dan-Eric Nilsson, Invertebrate Vision (2006). Information on the structure and photoreception mechanisms of the human eye is provided in Clyde W. Oyster, The Human Eye: Structure and Function (1999).

Evolution and ecology of vision
The eye in the context of the evolution of organisms is covered in intriguing detail in Richard Dawkins, Climbing Mount Improbable (1996). Works providing information on the ecology and adaptive mechanisms of photoreception in the eye are John N. Lythgoe, The Ecology of Vision (1979); and Simon N. Archer et al. (eds.), The Adaptive Mechanisms in the Ecology of Vision (1998).

Vision and the brain
The mechanisms of photoreception and central processing of visual information are discussed in David H. Hubel, Eye, Brain, and Vision (1995); A. David Milner and Melvyn A. Goodale, The Visual Brain in Action, 2nd ed. (2006); and Semir Zeki, A Vision of the Brain (1993).Michael Land

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • photoreception — noun Date: 1902 perception of waves in the range of visible light; specifically vision • photoreceptive adjective …   New Collegiate Dictionary

  • photoreception — noun the absorption of light energy by plants and animals, and its utilization for biologically important purposes …   Wiktionary

  • photoreception — pho·to·re·cep·tion ri sep shən n perception of waves in the range of visible light specif VISION (1) pho·to·re·cep·tive sep tiv adj * * * pho·to·re·cep·tion (fo″to re sepґshən) [photo + re + ceptor] the process of detecting radiant… …   Medical dictionary

  • photoreception — n. perception of light …   English contemporary dictionary

  • photoreception — pho·to·reception …   English syllables

  • photoreception — pho•to•re•cep•tion [[t]ˌfoʊ toʊ rɪˈsɛp ʃən[/t]] n. phl the physiological perception of light • Etymology: 1905–10 pho to•re•cep′tive, adj …   From formal English to slang

  • photoreception — …   Useful english dictionary

  • Cryptochrome — 1 (photolyase like) Crystal structure of the PHR domain of cryptochrome 1 from Arabidopsis thaliana.[1] …   Wikipedia

  • VISION (physiologie) — L’univers sensoriel de l’Homme est essentiellement visuel: telle est probablement la raison pour laquelle la physiologie de la vision occupe une place si considérable dans l’ensemble de la physiologie sensorielle. Depuis les années trente, qui… …   Encyclopédie Universelle

  • Evolution de l'oeil — Histoire évolutive de l œil L œil est un système sensoriel utilisé dans l ensemble du règne animal. En effet, la lumière donnant un grand nombre d informations sur l environnement, elle est, et a été, une force sélective majeure, qui a conduit à… …   Wikipédia en Français

Share the article and excerpts

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