Difference between revisions of "Rod cell" - New World Encyclopedia

From New World Encyclopedia
m (Protected "Rod cell": Copyedited [edit=sysop:move=sysop])
(No difference)

Revision as of 18:55, 19 May 2008

Rod cell
Rod cell - Cross section of the retina. Rods are visible at far right.
Cross section of the retina. Rods are visible at far right.
Location Retina
Function Low light photoreceptor
Morphology rod shaped
Presynaptic connections None
Postsynaptic connections Bipolar Cells and Horizontal cells

A rod cell, or rod, is any of the generally cylindrically or rod-shaped photoreceptor cells in the retina of the eye that are sensitive to dim light and lack the visual acuity and color-distinguishing ability of the other type of photoreceptor, cone cells. Since they can function in less intense light than cone cells, rods are responsible for night vision in humans and predominate in nocturnal vertebrates. Named for their cylindrical shape, these photosensitive cells are concentrated at the outer edges of the retina and are used in peripheral vision. There are about 120 million rod cells in the human retina.

The human visual system of rods and cones is a complementary one, allowing one to see in both low light conditions (rods) and to see a diversity of colors in brighter light (cones). While cones allow humans to experience the great beauty that color adds to one's environment, rods allow one to see in dim light, opening up the experience of one's environment. Thus, one can take a walk in moonlight or explore a cave or see distant stars at night.

Overview

The retina contains two forms of photosensitive cells—rods and cones. Though structurally and metabolically similar, their function is quite different. Rod cells are highly sensitive to light, allowing them to respond in dim light and dark conditions. These are the cells that allow humans and other animals to see by moonlight, or with very little available light (as in a dark room). However, they do not distinguish between colors, and have low visual acuity (measure of detail). This is why the darker conditions become, the less color objects seem to have. Cone cells, conversely, need high light intensities to respond and have high visual acuity. Different cone cells respond to different colors (wavelengths of light), which allows an organism to see color.

Rods and cones are both photosensitive, but respond differently to different frequencies of light. They both contain different pigmented photoreceptor proteins. Rod cells contain the protein rhodopsin and cone cells contain different proteins for each color-range. The process through which these proteins work is quite similar—upon being subjected to electromagnetic radiation of a particular wavelength and intensity, the protein breaks down into two constituent products. Rhodopsin, of rods, breaks down into opsin and retinal; iodopsin of cones breaks down into photopsin and retinal. The opsin in both opens ion channels on the cell membrane which leads to the generation of an action potential (an impulse which will eventually get to the visual cortex in the brain).

This is the reason why cones and rods enable organisms to see in dark and light conditions—each of the photoreceptor proteins requires a different light intensity to break down into the constituent products. Further, synaptic convergence means that several rod cells are connected to a single bipolar cell, which then connects to a single ganglion cell and information is relayed to the visual cortex. On the other hand, a single cone cell is connected to a single bipolar cell. Thus, action potentials from rods share neurons, whereas those from cones are given their own. This results in the high visual acuity, or the high ability to distinguish between detail, of cone cells and not rods. If a ray of light were to reach just one rod cell this may not be enough to stimulate an action potential. Because several "converge" onto a bipolar cell, enough transmitter molecules reach the synapse of the bipolar cell to attain the threshold level to generate an action potential.

Oyster (1999) cites evidence for an average of about 90 million rod cells and 4.5 million cone cells in the human retina.

Description and sensitivity

Like cones, rod cells have a synaptic terminal, an inner segment, and an outer segment. The synaptic terminal forms a synapse with another neuron, for example a bipolar cell. The inner and outer segments are connected by a cilium (Kandel et al. 2000). The inner segment contains organelles and the cell's nucleus, while the outer segment, which is pointed toward the front of the eye, contains the light-absorbing materials (Kandel et al. 2000).

Like cones, the outer segments of rods have invaginations of their cell membranes that create stacks of membranous disks. Photopigments exist as transmembrane proteins within these disks, which provide more surface area for light to affect the pigments. In cones, these disks are attached to the outer membrane, whereas they are pinched off and exist separately in rods. Neither rods nor cones divide, but their membranous disks wear out and are worn off at the end of the outer segment, to be consumed and recycled by phagocytic cells.

A rod cell is sensitive enough to respond to a single photon of light, and is about 100 times more sensitive to a single photon than cones. Since rods require less light to function than cones, they are therefore the primary source of visual information at night (scotopic vision). Cone cells, on the other hand, require tens to hundreds of photons to become activated. Additionally, multiple rod cells converge on a single interneuron, collecting and amplifying the signals. However, this convergence comes at a cost to visual acuity (or Image resolution) since the pooled information from multiple cells is less distinct than it would be if the visual system received information from each rod cell individually. The convergence of rod cells also tends to make peripheral vision very sensitive to movement, and is responsible for the phenomenon of an individual seeing something vague occur out of the corner of his or her eye.

Rod cells also respond more slowly to light than cones do, so stimuli they receive are added over about 100 milliseconds. While this makes rods more sensitive to smaller amounts of light, it also means that their ability to sense temporal changes, such as quickly changing images, is less accurate than that of cones (Kandel et al. 2000). However, if multiple flashes of sub-threshold light occurs during the 100 millisecond period, the energy of the flashes of light would summate to produce a light that will reach threshold and send a signal to the brain.

Experiments by George Wald and others showed that rods are more sensitive to the blue area of the spectrum, and are completely insensitive to wavelengths above about 640 nm (red). This fact is responsible for the Purkinje effect, in which blue colors appear more intense relative to reds in darker light, when rods take over as the cells responsible for vision.

In humans, the fovea, directly behind the lens, consists of mostly densely-packed cone cells. Its requirement for high intensity light does cause problems for astronomers, as they cannot see dim stars, or other objects, using central vision because the light from these is not enough to stimulate cone cells. Because cone cells are all that exist directly in the fovea, astronomers have to look at stars through the "corner of their eyes" (averted vision) where rods also exist, and where the light can stimulate cells, allowing the individual to observe distant stars. Because they have only one type of light sensitive pigment, rather than the three types that human cone cells have, rods have little, if any, role in color vision.

Response to light

Activation of a photoreceptor cell is actually a hyperpolarization (inhibition) of the cell. When they are not being stimulated, such as in the dark, rods and cones depolarize and release a neurotransmitter spontaneously. This neurotransmitter hyperpolarizes the bipolar cell. Bipolar cells exist between photoreceptors and ganglion cells and act to transmit signals from the photoreceptors to the ganglion cells. As a result of the bipolar cell being hyperpolarized, it does not release its transmitter at the bipolar-ganglion synapse and the synapse is not excited.

Activation of photopigments by light sends a signal by hyperpolarizing the rod cell, leading to the rod cell not sending its neurotransmitter, which leads to the bipolar cell then releasing its transmitter at the bipolar-ganglion synapse and exciting the synapse.

Depolarization of rod cells (causing release of their neurotransmitter) occurs because in the dark, cells have a relatively high concentration of cyclic guanosine 3'-5' monophosphate (cGMP), which opens ion channels (largely sodium channels, though calcium can enter through these channels as well). The positive charges of the ions that enter the cell down its electrochemical gradient change the cell's membrane potential, cause depolarization, and lead to the release of the neurotransmitter glutamate. Glutamate can depolarize some neurons and hyperpolarize others, allowing photoreceptors to interact in an antagonistic manner.

When light hits photoreceptive pigments within the photoreceptor cell, the pigment changes shape. The pigment, called rhodopsin (iodopsin is found in cone cells) consists of a large protein called opsin (situated in the plasma membrane), attached to which is a covalently-bound prosthetic group: an organic molecule called retinal (a derivative of vitamin A). The retinal exists in the 11-cis-retinal form when in the dark, and stimulation by light causes its structure to change to all-trans-retinal. This structural change causes it to activate a regulatory protein called transducin, which leads to the activation of cGMP phosphodiesterase, which breaks cGMP down into 5'-GMP. Reduction in cGMP allows the ion channels to close, preventing the influx of positive ions, hyperpolarizing the cell, and stopping the release of neurotransmitters (Kandel et al., 2000). Though cone cells primarily use the transmitter substance acetyl choline, rod cells use a variety. The entire process by which light initiates a sensory response is called visual phototransduction.

Activation of a single molecule of rhodopsin, the photosensitive pigment in rods, can lead to a large reaction in the cell because the signal is amplified. Once activated, rhodopsin can activate hundreds of transducin molecules, each of which in turn activate a phosphodiesterase molecule, which can break down over a thousand cGMP molecules per second (Kandel et al. 2000). Thus, rods can have a large response to a small amount of light.

As the retinal component of rhodopsin is derived from vitamin A, a deficiency of vitamin A causes a deficit in the pigment needed by rod cells. Consequently, fewer rod cells are able to sufficiently respond in darker conditions, and as the cone cells are poorly adapted for sight in the dark, blindness can result. This is night-blindness.

Revert to the resting state

Rods make use of three inhibitory mechanisms (negative feedback mechanisms) to allow a rapid revert to the resting state after a flash of light.

Firstly, there exists a rhodopsin kinase (RK), which would phosphorylate the cytosolic tail of the activated rhodopsin on the multiple serines, partially inhibiting the activation of transducin. Also, an inhibitory protein, arrestin, then binds to the phosphorylated rhodopsins to further inhibit the rhodopsin's activity.

While arrestin shuts off rhodopsin, an RGS protein (functioning as a GTPase-activiating proteins(GAPs)) drives the transducin (G-protein) into an "off" state by increasing the rate of hydrolysis of the bounded GTP to GDP.

Also, as the cGMP sensitive channels allow not only the influx of sodium ions, but also calcium ions, with the decrease in concentration of cGMP, cGMP sensitive channels are then closed and reducing the normal influx of calcium ions. The decrease in the concentration of calcium ions stimulates the calcium ion-sensitive proteins, which would then activiate the guanylyl cyclase to replenish the cGMP, rapidly restoring its original concentration. The restoration opens the cGMP sensitive channels and causes a depolarization of the plasma membrane (Alberts et al. 2008).

Desensitization

When the rods are exposed to a high concentration of photons for a prolonged period, they become desensitized (adapted) to the environment.

As rhodopsin is phosphorylated by rhodopsin kinase (a member of the GPCR kinases(GRKs)), it binds with high affinity to the arrestin. The bound arrestin can contribute to the densitization process in at least two ways. First, it prevents the interaction between the G-protein and the activated receptor. Second, it serves as an adaptor protein to aid the receptor to the clathrin-dependent endocytosis machinery (to induce receptor-mediated endocytosis) (Alberts et al. 2008).

Table

Comparison of rod and cone cells, from Kandel et al. (2000).

Rods Cones
Used for night vision Used for day vision
Highly sensitive to light; sensitive to scattered light (they have more pigment than cones) At least a tenth of the rods' light sensitivity; sensitive only to direct light
Loss causes night blindness Loss constitutes legal blindness
Low spatial resolution with higher noise High spatial resolution with lower noise
Not present in the fovea Concentrated in the fovea
Slower response to light; rods need to be exposed to light over time Quicker response to light; can perceive more rapid changes in stimuli
Stacks of membrane-enclosed disks are unattached to the cell membrane Disks are attached to the outer membrane
22 times as numerous as cones in the retina
One type of photosensitive pigment (monochromatic stimulus) Three types of photosensitive pigment in humans (trichromatic stimulus)
Confer achromatic vision, with more emphasis on detecting motion Confer colour vision, with more emphasis on detecting fine details

References
ISBN links support NWE through referral fees

  • Alberts, B., A. Johnson, J. Lewis, M. Raff, K. Roberts, and P. Walter. 2008. Molecular Biology of The Cell, 5th ed. Garland Science. ISBN 9780815341116.
  • Kandel, E. R., J. H. Schwartz, and T. M. Jessell. 2000. Principles of Neural Science, 4th ed. McGraw-Hill, New York. ISBN 0071120009.
  • Osterberg, G. 1935. Topography of the layer of rods and cones in the human retina. Acta Ophthalmol. Suppl. 6: 1–103.
  • Oyster, C. W. 1999. The Human Eye: Structure and Function. Sunderland, Mass: Sinauer Associates. ISBN 0878936459.

Credits

New World Encyclopedia writers and editors rewrote and completed the Wikipedia article in accordance with New World Encyclopedia standards. This article abides by terms of the Creative Commons CC-by-sa 3.0 License (CC-by-sa), which may be used and disseminated with proper attribution. Credit is due under the terms of this license that can reference both the New World Encyclopedia contributors and the selfless volunteer contributors of the Wikimedia Foundation. To cite this article click here for a list of acceptable citing formats.The history of earlier contributions by wikipedians is accessible to researchers here:

The history of this article since it was imported to New World Encyclopedia:

Note: Some restrictions may apply to use of individual images which are separately licensed.