nervous system

nervous system
Anat., Zool.
1. the system of nerves and nerve centers in an animal or human, including the brain, spinal cord, nerves, and ganglia.
[1730-40]

* * *

System of specialized cells (neurons, or nerve cells) that conduct stimuli from a sensory receptor through a neuron network to the site (e.g., a gland or muscle) where the response occurs.

In humans, it consists of the central and peripheral nervous systems, the former consisting of the brain and spinal cord and the latter of the nerves, which carry impulses to and from the central nervous system. The cranial nerves handle head and neck sensory and motor activities, except the vagus nerve, which conducts signals to visceral organs. Each spinal nerve is attached to the spinal cord by a sensory and a motor root. These exit between the vertebrae and merge to form a large mixed nerve, which branches to supply a defined area of the body. Disorders include amyotrophic lateral sclerosis, chorea, epilepsy, myasthenia gravis, neural tube defect, parkinsonism, and poliomyelitis. Effects of disorders range from transient tics and minor personality changes to major personality disruptions, seizures, paralysis, and death.

* * *

Introduction

      organized group of cells specialized for the conduction of electrochemical stimuli from sensory receptors through a network to the site at which a response occurs.

 All living organisms are able to detect changes within themselves and in their environments. Changes in the external environment include those of light, temperature, sound, motion, and odour, while changes in the internal environment include those in the position of the head and limbs as well as in the internal organs. Once detected, these internal and external changes must be analyzed and acted upon in order to survive. As life on Earth evolved and the environment became more complex, the survival of organisms depended upon how well they could respond to changes in their surroundings. One factor necessary for survival was a speedy reaction or response. Since communication from one cell to another by chemical means was too slow to be adequate for survival, a system evolved that allowed for faster reaction. That system was the nervous system, which is based upon the almost instantaneous transmission of electrical impulses from one region of the body to another along specialized cells called neurons (neuron).

  Nervous systems are of two general types, diffuse and centralized. In the diffuse type of system, found in lower invertebrates (invertebrate), there is no brain, and neurons are distributed throughout the organism in a netlike pattern. In the centralized systems of higher invertebrates and vertebrates (vertebrate), a portion of the nervous system has a dominant role in coordinating information and directing responses. This centralization reaches its culmination in vertebrates, which have a well-developed brain and spinal cord. Impulses are carried to and from the brain and spinal cord by nerve fibres that make up the peripheral nervous system.

      This article begins with a discussion of the general features of nervous systems—that is, their function of responding to stimuli and the rather uniform electrochemical processes by which they generate a response. Following that is a discussion of the various types of nervous systems, from the simplest to the most complex.

Solomon D. Erulkar

Form and function of nervous systems

Stimulus-response coordination
      The simplest type of response is a direct one-to-one stimulus-response reaction. A change in the environment is the stimulus; the reaction of the organism to it is the response. In single-celled organisms, the response is the result of a property of the cell fluid called irritability. In simple organisms, such as algae, protozoans (protozoan), and fungi (fungus), a response in which the organism moves toward or away from the stimulus is called taxis. In larger and more complicated organisms—those in which response involves the synchronization and integration of events in different parts of the body—a control mechanism, or controller, is located between the stimulus and the response. In multicellular organisms, this controller consists of two basic mechanisms by which integration is achieved—chemical regulation and nervous regulation.

      In chemical regulation, substances called hormones (hormone) are produced by well-defined groups of cells and are either diffused or carried by the blood to other areas of the body where they act on target cells and influence metabolism or induce synthesis of other substances. The changes resulting from hormonal action are expressed in the organism as influences on, or alterations in, form, growth, reproduction, and behaviour.

      Plants respond to a variety of external stimuli by utilizing hormones as controllers in a stimulus-response system. Directional responses of movement are known as tropisms (tropism) and are positive when the movement is toward the stimulus and negative when it is away from the stimulus. When a seed germinates, the growing stem turns upward toward the light, and the roots turn downward away from the light. Thus, the stem shows positive phototropism and negative geotropism, while the roots show negative phototropism and positive geotropism. In this example, light and gravity are the stimuli, and directional growth is the response. The controllers are certain hormones synthesized by cells in the tips of the plant stems. These hormones, known as auxins (auxin), diffuse through the tissues beneath the stem tip and concentrate toward the shaded side, causing elongation of these cells and, thus, a bending of the tip toward the light. The end result is the maintenance of the plant in an optimal condition with respect to light.

      In animals, in addition to chemical regulation via the endocrine system, there is another integrative system called the nervous system. A nervous system can be defined as an organized group of cells, called neurons, specialized for the conduction of an impulse—an excited state—from a sensory receptor through a nerve network to an effector, the site at which the response occurs.

      Organisms that possess a nervous system are capable of much more complex behaviour than are organisms that do not. The nervous system, specialized for the conduction of impulses, allows rapid responses to environmental stimuli. Many responses mediated by the nervous system are directed toward preserving the status quo, or homeostasis, of the animal. Stimuli that tend to displace or disrupt some part of the organism call forth a response that results in reduction of the adverse effects and a return to a more normal condition. Organisms with a nervous system are also capable of a second group of functions that initiate a variety of behaviour patterns. Animals may go through periods of exploratory or appetitive behaviour, nest building, and migration. Although these activities are beneficial to the survival of the species, they are not always performed by the individual in response to an individual need or stimulus. Finally, learned behaviour can be superimposed on both the homeostatic and initiating functions of the nervous system.

Intracellular systems
      All living cells have the property of irritability, or responsiveness to environmental stimuli, which can affect the cell in different ways, producing, for example, electrical, chemical, or mechanical changes. These changes are expressed as a response, which may be the release of secretory products by gland cells, the contraction of muscle cells, the bending of a plant-stem cell, or the beating of whiplike “hairs,” or cilia (cilium), by ciliated cells.

      The responsiveness of a single cell can be illustrated by the behaviour of the relatively simple amoeba. Unlike some other protozoans, an amoeba lacks highly developed structures that function in the reception of stimuli and in the production or conduction of a response. The amoeba behaves as though it had a nervous system, however, because the general responsiveness of its cytoplasm serves the functions of a nervous system. An excitation produced by a stimulus is conducted to other parts of the cell and evokes a response by the animal. An amoeba will move to a region of a certain level of light. It will be attracted by chemicals given off by foods and exhibit a feeding response. It will also withdraw from a region with noxious chemicals and exhibit an avoidance reaction upon contacting other objects.

organelle systems
      In more-complex protozoans, specialized cellular structures, or organelles, serve as receptors of stimulus and as effectors of response. Receptors include stiff sensory bristles in ciliates (ciliate) and the light-sensitive eyespots of flagellates (flagellate). Effectors include cilia (slender, hairlike projections from the cell surface), flagella (flagellum) (elongated, whiplike cilia), and other organelles associated with drawing in food or with locomotion. Protozoans (protozoan) also have subcellular cytoplasmic filaments that, like muscle tissue, are contractile. The vigorous contraction of the protozoan Vorticella, for example, is the result of contraction of a threadlike structure called a myoneme in the stalk.

      Although protozoans clearly have specialized receptors and effectors, it is not certain that there are special conducting systems between the two. In a ciliate such as Paramecium, the beating of the cilia—which propels it along—is not random, but coordinated. Beating of the cilia begins at one end of the organism and moves in regularly spaced waves to the other end, suggesting that coordinating influences are conducted longitudinally. A system of fibrils connecting the bodies in which the cilia are rooted may provide conducting paths for the waves, but coordination of the cilia may also take place without such a system. Each cilium may respond to a stimulus carried over the cell surface from an adjacent cilium—in which case, coordination would be the result of a chain reaction from cilium to cilium.

      The best evidence that formed structures are responsible for coordination comes from another ciliate, Euplotes, which has a specialized band of ciliary rows (membranelles) and widely separated tufts of cilia (cirri). By means of the coordinated action of these structures, Euplotes is capable of several complicated movements in addition to swimming (e.g., turning sharply, moving backward, spinning). The five cirri at the rear of the organism are connected to the anterior end in an area known as the motorium. The fibres of the motorium apparently provide coordination between the cirri and the membranelles. The membranelles, cirri, and motorium constitute a neuromotor system.

Nervous systems
      The basic pattern of stimulus-response coordination in animals is an organization of receptor, adjustor, and effector units. External stimuli are received by the receptor cells, which, in most cases, are neurons. (In a few instances, a receptor is a non-nervous sensory epithelial cell, such as a hair cell of the inner ear or a taste cell, which stimulates adjacent neurons.) The stimulus is modified, or transduced, into an electrical impulse in the receptor neuron. This incoming excitation, or afferent impulse, then passes along an extension, or axon, of the receptor to an adjustor, called an interneuron. (All neurons are capable of conducting an impulse, which is a brief change in the electrical charge on the cell membrane. Such an impulse can be transmitted, without loss in strength, many times along an axon until the message, or input, reaches another neuron, which in turn is excited.) The interneuron-adjustor selects, interprets, or modifies the input from the receptor and sends an outgoing, or efferent, impulse to an efferent neuron, such as a motor neuron. The efferent neuron, in turn, makes contact with an effector such as a muscle or gland, which produces a response.

      In the simplest arrangement, the receptor-adjustor-effector units form a functional group known as the reflex arc (reflex). Sensory cells carry afferent impulses to a central interneuron, which makes contact with a motor neuron. The motor neuron carries efferent impulses to the effector, which produces the response. Three types of neurons are involved in this reflex arc, but a two-neuron arc, in which the receptor makes contact directly with the motor neuron, also occurs. In a two-neuron arc, simple reflexes are prompt, short-lived, and automatic and involve only a part of the body. Examples of simple reflexes are the contraction of a muscle in response to stretch, the blink of the eye when the cornea is touched, and salivation at the sight of food. Reflexes of this type are usually involved in maintaining homeostasis.

      The differences between simple and complex nervous systems lie not in the basic units but in their arrangement. In higher nervous systems, there are more interneurons concentrated in the central nervous system ( brain and spinal cord) that mediate the impulses between afferent and efferent neurons. Sensory impulses from particular receptors travel through specific neuronal pathways to the central nervous system. Within the central nervous system, though, the impulse can travel through multiple pathways formed by numerous neurons. Theoretically, the impulse can be distributed to any of the efferent motor neurons and produce a response in any of the effectors. It is also possible for many kinds of stimuli to produce the same response.

      As a result of the integrative action of the interneuron, the behaviour of the organism is more than the simple sum of its reflexes; it is an integrated whole that exhibits coordination between many individual reflexes. Reflexes can occur in a complicated sequence producing elaborate behaviour patterns. Behaviour in such cases is characterized not by inherited, stereotyped responses but by flexibility and adaptability to circumstances. Many automatic, unconditioned reflexes can be modified by or adapted to new stimuli. The experiments of the Russian physiologist Ivan Petrovich Pavlov (Pavlov, Ivan Petrovich), for example, showed that if an animal salivates at the sight of food while another stimulus, such as the sound of a bell, occurs simultaneously, the sound alone can induce salivation after several trials. This response, known as a conditioned reflex, is a form of learning. The behaviour of the animal is no longer limited by fixed, inherited reflex arcs but can be modified by experience and exposure to an unlimited number of stimuli. The most evolved nervous systems are capable of even higher associative functions such as thinking and memory. The complex manipulation of the signals necessary for these functions depends to a great extent on the number and intricacy of the arrangement of interneurons.

Thomas L. Lentz Solomon D. Erulkar

The nerve cell (neuron)
      The watershed of all studies of the nervous system was an observation made in 1889 by the Spanish scientist Santiago Ramón y Cajal (Ramón y Cajal, Santiago), who reported that the nervous system is composed of individual units that are structurally independent of one another and whose internal contents do not come into direct contact. According to his hypothesis, now known as the neuron theory, each nerve cell communicates with others through contiguity rather than continuity. That is, communication between adjacent but separate cells must take place across the space and barriers separating them. It has since been proved that Cajal's theory is not universally true, but his central idea—that communication in the nervous system is largely communication between independent nerve cells—has remained an accurate guiding principle for all further study.

      There are two basic cell types within the nervous system: neurons and neuroglial cells.

The neuron
 In the human brain there are approximately 10 billion neurons. Each neuron has its own identity, expressed by its interactions with other neurons and by its secretions; each also has its own function, depending on its intrinsic properties and location as well as its inputs from other select groups of neurons, its capacity to integrate those inputs, and its ability to transmit the information to another select group of neurons.

 With few exceptions, most neurons consist of three distinct regions, as shown in the diagram—>: (1) the cell body, or soma; (2) the nerve fibre, or axon; and (3) the receiving processes, or dendrites.

 The neuron is bound by a plasma membrane, a structure so thin that its fine detail can be revealed only by high-resolution electron microscopy, as shown in the photograph—>. About half of the membrane is the lipid bilayer, two sheets of mainly phospholipids (phospholipid) with a space between. One end of a phospholipid molecule is hydrophilic, or water attaching, and the other end is hydrophobic, or water repelling. The bilayer structure results when the hydrophilic ends of the phospholipid molecules in each sheet turn toward the watery mediums of both the cell interior and the extracellular environment, while the hydrophobic ends of the molecules turn in toward the space between the sheets. These lipid layers are not rigid structures; the loosely bonded phospholipid molecules can move laterally across the surfaces of the membrane, and the interior is in a highly liquid state.

      Embedded within the lipid bilayer are proteins, which also float in the liquid environment of the membrane. These include glycoproteins containing polysaccharide chains, which function, along with other carbohydrates, as adhesion sites and recognition sites for attachment and chemical interaction with other neurons. The proteins provide another basic and crucial function: those which penetrate the membrane can exist in more than one conformational state, or molecular shape, forming channels that allow ions to pass between the extracellular fluid and the cytoplasm, or internal contents of the cell. In other conformational states, they can block the passage of ions. This action is the fundamental mechanism that determines the excitability and pattern of electrical activity of the neuron.

      A complex system of proteinaceous intracellular filaments is linked to the membrane proteins. This cytoskeleton includes thin neurofilaments containing actin, thick neurofilaments similar to myosin, and microtubules (microtubule) composed of tubulin. The filaments are probably involved with movement and translocation of the membrane proteins, while microtubules may anchor the proteins to the cytoplasm.

      Each neuron contains a nucleus defining the location of the soma. The nucleus is surrounded by a double membrane, called the nuclear envelope, that fuses at intervals to form pores allowing molecular communication with the cytoplasm. Within the nucleus are the chromosomes (chromosome), the genetic material of the cell, through which the nucleus controls the synthesis of proteins and the growth and differentiation of the cell into its final form. Proteins synthesized in the neuron include enzymes, receptors, hormones, and structural proteins for the cytoskeleton.

Organelles (organelle)
      The endoplasmic reticulum (ER) is a widely spread membrane system within the neuron that is continuous with the nuclear envelope. It consists of series of tubules, flattened sacs called cisternae, and membrane-bound spheres called vesicles. There are two types of ER. The rough endoplasmic reticulum (RER) has rows of knobs called ribosomes (ribosome) on its surface. Ribosomes synthesize proteins that, for the most part, are transported out of the cell. The RER is found only in the soma. The smooth endoplasmic reticulum (SER) consists of a network of tubules in the soma that connects the RER with the Golgi apparatus. The tubules can also enter the axon at its initial segment and extend to the axon terminals.

      The Golgi apparatus is a complex of flattened cisternae arranged in closely packed rows. Located close to and around the nucleus, it receives proteins synthesized in the RER and transferred to it via the SER. At the Golgi apparatus, the proteins are attached to carbohydrates. The glycoproteins so formed are packaged into vesicles that leave the complex to be incorporated into the cell membrane.

 The axon arises from the soma at a region called the axon hillock, or initial segment. This is the region where the plasma membrane generates nerve impulses; the axon conducts these impulses away from the soma or dendrites toward other neurons. Large axons acquire an insulating myelin sheath and are known as myelinated, or medullated, fibres. myelin is composed of 80 percent lipid and 20 percent protein; cholesterol is one of the major lipids, along with variable amounts of cerebrosides and phospholipids. Concentric layers of these lipids separated by thin layers of protein give rise to a high-resistance, low-capacitance electrical insulator interrupted at intervals by gaps called nodes of Ranvier, where the nerve membrane is exposed to the external environment. (See the diagram—>.) In the central nervous system, the myelin sheath is formed from glial cells called oligodendrocytes, while in peripheral nerves it is formed from Schwann cells see below The neuroglia (nervous system).

      While the axon mainly conducts nerve impulses from the soma to the terminal, the terminal itself secretes chemical substances called neurotransmitters (neurotransmitter). The synthesis of these substances can occur in the terminal itself, but the synthesizing enzymes are formed by ribosomes in the soma and must be transported down the axon to the terminal. This process is known as axoplasmic flow; it occurs in both directions along the axon and may be facilitated by microtubules.

      At the terminal of the axon, and sometimes along its length, are specialized structures that form junctions with other neurons and with muscle cells. These junctions are called synapses (synapse). Presynaptic terminals, when seen by light microscope, look like small knobs and contain many organelles. The most numerous of these are synaptic vesicles, which, filled with neurotransmitters, are often clumped in areas of the terminal membrane that appear to be thickened. The thickened areas are called presynaptic dense projections, or active zones.

      The presynaptic terminal is unmyelinated and is separated from the neuron or muscle cell onto which it impinges by a gap called the synaptic cleft, across which neurotransmitters diffuse when released from the vesicles. In nerve-muscle junctions the synaptic cleft contains a structure called the basal lamina, which holds an enzyme that destroys neurotransmitters and thus regulates the amount that reaches the postsynaptic receptors on the receiving cell. Most knowledge of postsynaptic neurotransmitter receptors comes from studies of the receptor on muscle cells. This receptor, called the end plate, is a glycoprotein composed of five subunits. Other neurotransmitter receptors do not have the same structure, but they are all proteins and probably have subunits with a central channel that is activated by the neurotransmitter.

      While the chemically mediated synapse described above forms the majority of synapses in vertebrate nervous systems, there are other types of synapses in vertebrate brains and, in especially great numbers, in invertebrate and fish nervous systems. At these synapses there is no synaptic gap; instead, there are gap junctions, direct channels between neurons that establish a continuity between the cytoplasm of adjacent cells and a structural symmetry between the pre- and postsynaptic sites. Rapid neuronal communication at these junctions is probably electrical in nature. (For further discussion, see the section Transmission of information in the nervous system: Transmission at the synapse (nervous system).)

Dendrites
      Besides the axon, neurons have other branches called dendrites that are usually shorter than axons and are unmyelinated. Dendrites are thought to form receiving surfaces for synaptic input from other neurons. In many dendrites these surfaces are provided by specialized structures called dendritic spines, which, by providing discrete regions for the reception of nerve impulses, isolate changes in electrical current from the main dendritic trunk.

      The traditional view of dendritic function presumes that only axons conduct nerve impulses and only dendrites receive them, but dendrites can form synapses with dendrites and axons and even somata can receive impulses. Indeed, some neurons have no axon; in these cases nervous transmission is carried out by the dendrites.

The neuroglia
      Neurons form a minority of the cells in the nervous system. Exceeding them in number by at least 10 to 1 are neuroglial cells, which exist in the nervous systems of invertebrates as well as vertebrates. Neuroglia can be distinguished from neurons by their lack of axons and by the presence of only one type of process. In addition, they do not form synapses, and they retain the ability to divide throughout their life span. While neurons and neuroglia lie in close apposition to one another, there are no direct junctional specializations, such as gap junctions, between the two types. Gap junctions do exist between neuroglial cells.

Types of neuroglia
      Apart from conventional histological and electron-microscopic techniques, immunologic techniques are used to identify different neuroglial cell types. By staining the cells with antibodies that bind to specific protein constituents of different neuroglia, neurologists have been able to discern two (in some opinions, three) main groups of neuroglia: (1) astrocytes, subdivided into fibrous and protoplasmic types; (2) oligodendrocytes, subdivided into interfascicular and perineuronal types; and sometimes (3) microglia.

      Fibrous astrocytes are prevalent among myelinated nerve fibres in the white matter of the central nervous system. Organelles seen in the somata of neurons are also seen in astrocytes, but they appear to be much sparser. These cells are characterized by the presence of numerous fibrils in their cytoplasm. The main processes exit the cell in a radial direction (hence the name astrocyte, meaning “star-shaped cell”), forming expansions and end feet at the surfaces of vascular capillaries.

      Unlike fibrous astrocytes, protoplasmic astrocytes occur in the gray matter of the central nervous system. They have fewer fibrils within their cytoplasm, and cytoplasmic organelles are sparse, so that the somata are shaped by surrounding neurons and fibres. The processes of protoplasmic astrocytes also make contact with capillaries.

      Oligodendrocytes have few cytoplasmic fibrils but a well-developed Golgi apparatus. They can be distinguished from astrocytes by the greater density of both the cytoplasm and the nucleus, the absence of fibrils and of glycogen in the cytoplasm, and large numbers of microtubules in the processes. Interfascicular oligodendrocytes are aligned in rows between the nerve fibres of the white matter of the central nervous system. In gray matter, perineuronal oligodendrocytes are located in close proximity to the somata of neurons. In the peripheral nervous system, neuroglia that are equivalent to oligodendrocytes are called Schwann cells.

      Microglial cells are small cells with dark cytoplasm and a dark nucleus. It is uncertain whether they are merely damaged neuroglial cells or occur as a separate group in living tissue.

Neuroglial functions
      The term neuroglia means “nerve glue,” and these cells were originally thought to be structural supports for neurons. This is still thought to be plausible, but other functions of the neuroglia are now generally accepted. Oligodendrocytes and Schwann cells produce the myelin sheath around neuronal axons. Some constituent of the axonal surface stimulates Schwann cell proliferation; the type of axon determines whether there is loose or tight myelination of the axon. In tight myelination a glial cell wraps itself like a rolled sheet around a length of axon until the fibre is covered by several layers. Between segments of myelin wrapping are exposed sections called nodes of Ranvier, which are important in the transmission of nerve impulses. Myelinated nerve fibres are found only in vertebrates, leading biologists to conclude that they are an adaptation to transmission over relatively long distances.

      Another well-defined role of neuroglial cells is the repair of the central nervous system following injury. Astrocytes divide after injury to the nervous system and occupy the spaces left by injured neurons. The role of oligodendrocytes after injury is unclear, but they may proliferate and form myelin sheaths.

      When neurons of the peripheral nervous system are severed, they undergo a process of degeneration followed by regeneration; fibres regenerate in such a way that they return to their original target sites. Schwann cells that remain after nerve degeneration apparently determine the route. This route direction is also performed by astrocytes during development of the central nervous system. In the developing cerebral cortex and cerebellum of primates, astrocytes project long processes to certain locations, and neurons migrate along these processes to arrive at their final locations. Thus, neuronal organization is brought about to some extent by the neuroglia.

      Astrocytes are also thought to have high-affinity uptake systems for neurotransmitters such as glutamate and gamma-aminobutyric acid (GABA). This function is important in the modulation of synaptic transmission. Uptake systems tend to terminate neurotransmitter action at the synapses and may also act as storage systems for neurotransmitters when they are needed. For instance, when motor nerves are severed, nerve terminals degenerate and their original sites are occupied by Schwann cells. The synthesis of neurotransmitters by neurons apparently also requires the presence of neuroglial cells in the vicinity.

      Finally, the environment surrounding neurons in the brain consists of a network of very narrow extracellular clefts. In 1907 Italian biologist Emilio Lugaro suggested that neuroglial cells exchange substances with the extracellular fluid and in this way exert control on the neuronal environment. It has since been shown that glucose, amino acids, and ions—all of which influence neuronal function—are exchanged between the extracellular space and neuroglial cells. After high levels of neuronal activity, for instance, neuroglial cells can take up and spatially buffer potassium ions and thus maintain normal neuronal function.

Transmission of information in the nervous system
 In the nervous system of animals at all levels of the evolutionary scale, the signals containing information about a particular stimulus are electrical in nature. In the past the nerve fibre and its contents were compared to metal wire, while the membrane was compared to insulation around the wire. This comparison was erroneous for a number of reasons. First, the charge carriers in nerves are ions, not electrons, and the density of ions in the axon is much less than that of electrons in a metal wire. Second, the membrane of an axon is not a perfect insulator, so that the movement of current along the axon is not complete. Finally, nerve fibres are smaller than most wires, so that the currents they can carry are limited in amplitude.

The ionic (ion) basis of electrical signals
      Ions (ion) are atoms or groups of atoms that gain an electrical charge by losing or acquiring electrons. For example, in the reaction that forms salt from sodium and chlorine, each sodium atom donates an electron, which is negatively charged, to a chlorine atom. The result is sodium chloride (NaCl), composed of one positively charged sodium ion (Na+) and one negatively charged chloride ion (Cl). A positively charged ion is called a cation; a negatively charged ion, an anion. The electrical events that constitute signaling in the nervous system depend upon the distribution of such ions on either side of the nerve membrane. Underlying these distributions and their change are crucial physical-chemical principles.

diffusion of ions across a membrane

Uncharged molecules
      Molecules in solution move randomly; the energy for their movement is derived from thermal energy. When a permeable membrane (a membrane that allows molecules to cross it) divides a heavily concentrated solution from a less-concentrated solution, there occurs a diffusion of molecules through the membrane and down their concentration gradient—that is, from the fluid with the higher concentration to that with the lower concentration. The number of molecules moving per unit of time is called the flow rate, or flux rate. Diffusion continues until the concentrations on both sides of the membrane are equal. A condition of no net flux is then established with an equal, random diffusion of molecules in both directions. This is called the equilibrium (proprioception) state.

      A membrane with pores allowing passage of molecules of only a particular size is called a semipermeable membrane. The semipermeable membrane imposes a condition of restricted diffusion in which the flux rate of the diffusing material is controlled by the permeability of the membrane, which in turn is dictated by the size of the pores and is given a unit of measure called the permeability coefficient.

 The water molecule, like other molecules, diffuses down its concentration gradient. If a rigid vessel contains water on one side of a semipermeable membrane and an impermeant substance (a substance that cannot cross the membrane) on the other side, the water tends to cross the membrane, diluting the substance and increasing the hydrostatic pressure on the other side, as shown in the diagram—>. The pressure then will tend to push water back across the membrane in opposition to the net flux. When the pressure built up equals the diffusion of water in the opposite direction, no net flux occurs and equilibrium is established. The migration of water (or any solvent) across a membrane is called osmosis, and the pressure necessary to establish equilibrium is called osmotic pressure. Water moves from a region of low osmotic pressure to a region of high osmotic pressure.

      The above example refers to water in a container with rigid walls. The neuron, however, has somewhat flexible walls, so that as water enters it, the cell tends to increase in volume, or swell. There is a direct relation between osmotic pressure across the plasma membrane and the final volume of a cell at equilibrium, so that if the osmotic pressure of the cell exterior is halved, the equilibrium volume of the cell will be twice its original volume.

Ions
 When potassium chloride (KCl) is placed into solution, the elements separate into potassium cations (K+) and chloride anions (Cl). Ions follow much the same principles of diffusion as uncharged molecules. For example, if a highly concentrated solution of KCl is separated from a lower concentration by a semipermeable membrane—one that is permeable to cations only—then K+ from the higher concentration diffuses across the membrane, following its concentration gradient to the region of lower concentration (see the figure—>, part A). Cl, being blocked by the membrane, remains behind. At this point the diffusion of ions creates conditions quite different from the diffusion of uncharged molecules and water molecules. The movement of cations toward the less-concentrated solution creates a separation of electrical charge across the membrane—that is, a greater number of positively charged ions will have moved to the side with the less-concentrated solution of KCl, and the side of the membrane with the higher concentration will have a more negative charge. This separation of charge—actually a difference in electrical potential—is called the potential difference, and it is the starting point of all electrical events in nervous systems. When present in the plasma membrane of the neuron, the potential difference transforms the neuron into an electrolytic cell that is capable, upon stimulation, of generating and transmitting electrical impulses.

 Complicating the ionic diffusion process is the phenomenon that opposite charges attract. This means that, in the example above, some of the K+ diffusing across the membrane is electrostatically drawn back up its concentration gradient toward the Cl. This creates a situation, shown in part B of the diagram—>, in which two tendencies oppose each other: (1) the diffusing tendency of the cation down its concentration gradient; and (2) the electrostatic voltage force tending to draw the cation back. These two forces eventually reach a state of no net flux, when the number of cations that they draw in each direction across the membrane is equal. The system is then in electrochemical equilibrium. At equilibrium, one side of the membrane may still have a more negative charge than the other. The potential difference is then called the equilibrium potential. (It is also called the Nernst potential, after Walther Nernst (Nernst, Walther Hermann), the German physical chemist who, in the late 19th century, developed equations for calculating the electrical potential at which there is no longer a net flux of a specific ion across a membrane.)

      The law of electroneutrality states that in any single ionic solution a sum of negative electrical charges attracts an equal sum of positive electrical charges. If a solution of KCl is divided into two parts by a membrane that is permeable to both ions, then the equal concentration of KCl across the membrane preserves chemical equilibrium between the two sides, while the equal concentrations of K+ and Cl on each side preserve electroneutrality on each side as well. This equilibrium can be upset by the addition to side 1 of a large number of K+ and an equal charge of impermeant anions (that is, negatively charged ions other than Cl that cannot permeate the membrane). In this case electroneutrality on side 1 is preserved, since the sum of positive charges added to that side is equaled by the sum of added negative charges. However, chemical equilibrium between side 1 and side 2 is not preserved, since side 1 now has a greater concentration of ions than side 2. Therefore, K+ diffuses down its concentration gradient, crossing the membrane to side 2 while drawing Cl with it to preserve electroneutrality. Diffusion continues until a new state of electrochemical equilibrium is reached; this occurs when the ratio of K+ concentration (on side 2 to that on side 1) is equal to the ratio of Cl concentration (on side 1 to that on side 2). Stated mathematically, equilibrium is reached when

      This is known as the Donnan equilibrium, after Frederick George Donnan (Donnan, Frederick George), the British chemist who in 1911 first measured the changes brought about by adding an impermeant substance to one side of a divided solution at equilibrium.

      In the new state of equilibrium, both sides are electrically neutral, since the impermeant anions added to side 1 are equaled by the added K+, and the K+ that has diffused to side 2 is balanced by the Cl electrostatically drawn along with it. But the entire solution is not at osmotic equilibrium, because the larger amount of ions on side 1 tends to draw water from side 2. Osmotic equilibrium can be established by the addition of ions to side 2. Indeed, in the neuron, osmotic equilibrium is maintained partly because large amounts of K+ and impermeant anions inside the cell are balanced by large amounts of salt outside the cell.

The neuronal membrane
      The principles outlined above can be applied to the neuron and its ionic contents.

      The plasma membrane of the neuron is semipermeable, being highly permeable to K+ and slightly permeable to Cl and Na+. In the extracellular fluid, electroneutrality is preserved by a balance between a high concentration of Na+ on the one hand and a high concentration of Cl, as well as small quantities of impermeant anions such as bicarbonate, phosphate, and sulfate, on the other. In the cytoplasm, where K+ concentration is high, the concentration of Cl is much below that necessary to balance the sum of the positive charges. Electroneutrality is maintained there by negatively charged impermeant proteins and phosphates. Osmotic balance is maintained between the extracellular fluid and the cytoplasm by movement of water through the plasma membrane when the total concentration of particles on one side is not equal to that on the other.

      These three characteristics of the neuron—semipermeability of the membrane, osmotic balance, and electroneutrality on each side—create an equilibrium electrical potential at which the inside of the membrane is more negative than the outside. In most neurons this potential, called the membrane potential, is between −60 and −75 millivolts (mV; or thousandths of a volt; the minus sign indicates that the inner surface is negative). When the inside of the plasma membrane has a negative charge compared to the outside, the neuron is said to be polarized. Any change in membrane potential tending to make the inside even more negative is called hyperpolarization, while any change tending to make it less negative is called depolarization.

      As stated above, the Nernst potential is the potential difference that exists across a membrane when a particular ion, having reached equilibrium between the tendency to diffuse down its concentration gradient and the tendency to be drawn back by other ions, is in a state of no net flux. The plasma membrane of the neuron is highly permeable to K+, and in fact the recorded membrane potential of most neurons (−60 to −75 mV) is close to that predicted by the Nernst equation for K+. However, it is not exactly the same, because K+ is not the only ion affecting the membrane potential. The membrane is also slightly permeable to Na+ and Cl. The permeability to Na+ may be low, but the high concentration of this cation outside the cell and the slightly negative electric charge inside the cell tend to drive Na+ inward. This in turn causes the inside of the cell to depolarize, placing K+ out of equilibrium. As a consequence, K+ leaves the cell until an equilibrium state is reached in which the leak inward of Na+ is equaled by the leak outward of K+ and there is no net flux of ions. There is also a tendency for Cl to permeate the membrane, since that ion is at higher concentration outside the neuron than inside. Therefore, for an equilibrium state to be produced, the sum of all three net currents must equal zero.

      Given the concentrations of all three ions on each side of the membrane and the relative permeability of the membrane to each ion, researchers can calculate the combined effect of K+, Na+, and Cl on the membrane potential by using the so-called constant-field equation. This equation, by including relative permeability as an important factor, takes into account the phenomenon that the more permeable a membrane is to a particular ion, the greater is the influence of that ion on the membrane potential. The permeance of Na+, for example, is only a fraction of that of K+, and the permeance of Cl is lower yet; therefore, while the membrane potential is highly sensitive to changes in the concentration of K+, it is less affected by changes in Na+ and almost unaffected by changes in Cl.

Transmission in the neuron
      The discussion above demonstrates that the electrical potential existing in neurons is based on the distribution of ions across the plasma membrane and that this distribution comes about through permeation of the membrane. In fact, ions are almost always hydrated in the form of ion-water complexes, which have great difficulty in penetrating the hydrophobic lipid bilayer of the plasma membrane. Permeation actually occurs through protein structures embedded in the lipid bilayer and spanning the membrane from cytoplasm to extracellular fluid. These structures, sometimes pumping ions from one side to the other and sometimes merely providing channels through which diffusing ions can flow past the lipid molecules, maintain the ionic distribution that keeps the membrane polarized, and they also allow the abrupt changes in distribution that create nerve impulses. The protein structures are described in detail in the section Ion transport (nervous system). Following is a discussion of the electrical events, created by movement of ions, that lead to nervous transmission in the neuron.

 The electrical potential across the nerve membrane can be measured by placing one microelectrode within the neuron (usually in the soma) and a second microelectrode in the extracellular fluid. The microelectrode consists of a sharp-tipped glass capillary tube filled with conducting solution. Upon penetration of the neuron, the potential at the tip of the electrode becomes electrically negative in relation to the outside of the electrode. As described above and shown in the graph—>, the value of this negative charge is usually between −60 and −75 mV. This is the membrane potential of the neuron at rest (i.e., when it is not generating a nerve impulse), and for this reason it is called the resting potential.

      The resting potential is maintained by the sodium-potassium pump, which steadily discharges more positive charge from the cell than it allows in, and by the relatively high permeance of K+, which leaks out of the cell through its membrane channels faster than Na+ leaks in.

Localized potential
      When a physical stimulus, such as touch, taste, or colour, acts on a sensory receptor cell specifically designed to respond to that stimulus, then the energy of the stimulus (e.g., mechanical, chemical, light) is transduced, or transformed, into an electrical response. This response is called the receptor potential, a type of local potential that, when it reaches high enough amplitude, generates the nerve impulse. (Another type of local potential is the postsynaptic potential, which originates in chemical receptors at the synaptic cleft. See the section Transmission at the synapse: Chemical transmission (nervous system).)

      Sensory receptors transduce stimuli into electrical responses by activating ion channels in their membranes. For example, in the stretch receptors of neurons attached to muscle cells, the stretching action of the muscle is thought to put a mechanical stress on protein filaments of the cytoskeleton, which in turn alter the shape of ion channels, inducing them to open and allowing cations to diffuse into the cell. Receptor cells sensitive to chemical and light energy, on the other hand, activate ion channels through the second-messenger system (second messenger). In this system, stimulated receptor molecules on the surface of the cell membrane catalyze a series of enzymatic reactions within the cytoplasm; these reactions in turn release energy, which activates the ion channels.

      By permitting a flux of Na+ into the cell, the opening of ion channels slightly depolarizes the membrane. The extent to which the membrane is depolarized depends upon the extent to which the sodium channels are activated, and this in turn depends upon the strength and duration of the original stimulus at the receptor. If depolarization reaches what is called the threshold potential, it triggers the nerve impulse, or action potential see below. If it does not reach that amplitude, then the neuron remains at rest, and the local potential, through a process called passive spread, diffuses along the nerve fibre and back out through the membrane.

      When it is of the postsynaptic type, the local potential usually begins in the dendrites and spreads toward the soma and axon. It is at the initial segment of the axon where, if the local potential is of threshold amplitude, the nerve impulse is generated.

 Because it varies in amplitude, the local potential is said to be graded. The greater the influx of positive charge—and, consequently, depolarization of the membrane—the higher the grade. Beginning at the resting potential of a neuron (for instance, −75 mV), a local potential can be of any grade up to the threshold potential (for instance, −58 mV). At the threshold, voltage-dependent sodium channels become fully activated, and Na+ pours into the cell. Almost instantly, as shown in the graph—>, the membrane actually reverses polarity, and the inside acquires a positive charge in relation to the outside. This reverse polarity constitutes the nerve impulse. It is called the action potential because the positive charge then flows through the cytoplasm, activating sodium channels along the entire length of the nerve fibre. This series of activations, by propagating the action potential along the fibre with virtually no reduction in amplitude, gives the nerve impulse its regenerative property.

      Researchers call the nerve impulse an “all-or-none” (all-or-none law) reaction since there are no gradations between threshold potential and fully activated potential. The neuron is either at rest with a polarized membrane, or it is conducting a nerve impulse at reverse polarization. The reverse polarity of active neurons is measured at about +30 mV. This is close to the Nernst potential for Na+—that is, the membrane potential at which electrochemical equilibrium would be established if the membrane were completely permeable to Na+.

      As instantaneous as the opening of sodium channels at threshold potential is their closing at the peak of action potential. This is called sodium inactivation, and it is caused by gates within the channel that are sensitive to depolarization. Following sodium inactivation is the opening of potassium channels, which allows the diffusion of K+ out of the cell. The combined effect of sodium inactivation, which blocks the influx of cations, and potassium activation, which causes the efflux of other cations, is the immediate return of the cell membrane to a polarized state, with the inside negative in relation to the outside. After repolarization there is a period during which a second action potential cannot be initiated, no matter how large a stimulus current is applied to the neuron. This is called the absolute refractory period, and it is followed by a relative refractory period, during which another action potential can be generated, but only by a greater stimulus current than that originally needed. This period is followed by the return of the neuronal properties to the threshold levels originally required for the initiation of action potentials.

      The sequence of sodium activation–sodium inactivation–potassium activation creates a nerve impulse that is brief in duration, lasting only a few milliseconds, and that travels down the nerve fibre like a wave, the membrane depolarizing in front of the current and repolarizing behind. Because nerve impulses are not graded in amplitude, it is not the size of the action potential that is important in processing information within the nervous system; rather, it is the number and frequency with which the impulses are fired.

 As stated above, the action potential is propagated along the axon without any decrease in amplitude with distance. However, the velocity of conduction along the nerve fibre is dependent upon several factors. The first factor is the outside diameter of the nerve fibre. The fastest conduction velocity occurs in the largest diameter nerve fibres. This phenomenon has formed the basis for classifying mammalian nerve fibres into groups in order of decreasing diameter and decreasing conduction velocity. Another factor is the temperature of the nerve fibre. Conduction velocity increases at high temperatures and decreases at low temperatures. Indeed, nerve conduction can be blocked by the local application of cold to a nerve fibre. Conduction velocity is also affected by myelination of the nerve fibre. Since ions cannot cross the lipid content of the myelin sheath, they spread passively down the nerve fibre until reaching the unmyelinated nodes of Ranvier. The nodes of Ranvier are packed with a high concentration of ion channels, which, upon stimulation, propagate the nerve impulse to the next node. In this manner the action potential jumps quickly from node to node along the fibre in a process called saltatory conduction (from Latin saltare, “to jump”; see the figure—>).

Transmission at the synapse
      Once an action potential has been generated at the axon hillock, it is conducted along the length of the axon until it reaches the terminals, the fingerlike extensions of the neuron that are next to other neurons and muscle cells (see the section The nerve cell: The neuron (nervous system)). At this point there exist two methods for transmitting the action potential from one cell to the other. In electrical transmission, the ionic current flows directly through channels that couple the cells. In chemical transmission, a chemical substance called the neurotransmitter passes from one cell to the other, stimulating the second cell to generate its own action potential.

Electrical transmission
      This method of transmitting nerve impulses, while far less common than chemical transmission, occurs in the nervous systems of invertebrates and lower vertebrates, as well as in the central nervous systems of some mammals. Transmission takes place through gap junctions, which are protein channels that link the cellular contents of adjacent neurons. Direct diffusion of ions through these junctions allows the action potential to be transmitted with little or no delay or distortion, in effect synchronizing the response of an entire group of neurons. The channels often allow ions to diffuse in both directions, but some gated channels restrict transmission to only one direction.

Chemical transmission
      There are two classic preparations for the study of chemical transmission at the synapse. One is the vertebrate neuromuscular junction, and the other is the giant synapse of the squid Loligo. These sites have the advantage of being readily accessible for recording by electrodes—especially the squid synapse, which is large enough that electrodes can be inserted directly into the presynaptic terminal and postsynaptic fibre. In addition, only a single synapse is involved at these sites, whereas a single neuron of the central nervous system may have many synapses with many other neurons, each with a different neurotransmitter.

Neurotransmitter release
 Two factors are essential for the release of the neurotransmitter from the presynaptic terminal: (1) depolarization of the terminal and (2) the presence of calcium ions (Ca2+) in the extracellular fluid. The membrane of the presynaptic terminal contains voltage-dependent calcium channels that open when the membrane is depolarized by a nerve impulse, allowing Ca2+ to diffuse into the terminal along its concentration gradient. (See the figure—>.) Following the entrance of Ca2+ is the release of neurotransmitter.

      It is uncertain what happens in the time between Ca2+ entry and transmitter release. Ca2+ is known to be sequestered by certain organelles within the terminal, including the endoplasmic reticulum. The ions may attach to the membranes of synaptic vesicles, in some way facilitating their fusion with the nerve terminal membrane. They may also be removed from the terminal by exchange with extracellular Na+—a mechanism known to occur at some neuronal membranes. What is certain is that when the concentration of Ca2+ is increased within the terminal, the probability of transmitter release is also increased.

      Neurotransmitters are packed into small, membrane-bound synaptic vesicles. Each vesicle contains thousands of neurotransmitter molecules, and there are thousands of vesicles in each axon terminal. Once stimulated by Ca2+, the vesicles move through the cytoplasm and fuse their membranes with the plasma membrane of the terminal. The transmitter molecules are then expelled from the vesicles into the synaptic cleft. This expulsion process is called exocytosis. Vesicle membranes are then recovered from the plasma membrane through endocytosis. In this process the membranes are surrounded by a protein coat at the lateral margins of the synapse and are then transferred to cisternae, which form in the terminal during nerve stimulation. There the vesicles lose their coats, are probably refilled with neurotransmitter, and pinch off from the cisternae to become synaptic vesicles once more.

      Because the neurotransmitter chemicals are packed into separate, almost identically sized vesicles, their release into the synaptic cleft is said to be quantal—that is, they are expelled in parcels, each vesicle adding its contents incrementally to the contents released from other parcels. This quantal release of neurotransmitter has a critical influence on the electrical potential created in the postsynaptic membrane.

      After neurotransmitter is released from the presynaptic terminal, it diffuses across the synaptic cleft and binds to receptor proteins on the postsynaptic membrane. Some receptors are ion channels that open or close when their molecular configuration is altered by the binding action of the neurotransmitter. Others are membrane proteins that, upon activation, catalyze second-messenger reactions within the postsynaptic cell; these reactions in turn open or close the ion channels. Whether acting upon ion channels directly or indirectly, the neurotransmitter molecules cause a sudden change in the permeability of the membrane to specific ions. Exactly which ions now permeate the membrane vary according to the neurotransmitters and their receptors (see the section Neurotransmitters and neuromodulators (nervous system)), but the net result of a change in ion diffusion is a change in electrical potential across the membrane. This change is called the postsynaptic potential, or PSP. (In reference to the neuromuscular synapse, it is called the end-plate potential, or EPP.)

      The most common potential change is depolarization, caused by a net influx of cations (usually Na+). Because this infusion of positive charge brings the membrane potential toward the threshold at which the nerve impulse is generated, it is called an excitatory postsynaptic potential (EPSP). Other neurotransmitters stimulate a net efflux of positive charge (usually in the form of K+ diffusing out of the cell), leaving the inside of the membrane more negative. Because this hyperpolarization draws the membrane potential farther from the threshold, making it more difficult to generate a nerve impulse, it is called an inhibitory postsynaptic potential (IPSP). The interaction of competing EPSPs and IPSPs at the hundreds or even thousands of synapses on a single neuron determines whether the nerve impulse arriving at the presynaptic terminals will be regenerated in the postsynaptic membrane.

      The PSP is a type of local potential, having properties similar to the electrical potential set up at sensory receptor neurons (see the section Transmission in the neuron: Localized potential (nervous system)). Like the receptor potential, the PSP is a graded response, varying in amplitude according to the duration and amount of stimulation by neurotransmitters. At the neuromuscular junction, brief depolarizations measuring no more than one millivolt can be observed in the postsynaptic muscle membrane, even when it is at rest. These tiny electrical events, called miniature end-plate potentials (MEPPs), or miniature postsynaptic potentials (MPSPs), are caused by the random release of single quanta of neurotransmitter from a resting presynaptic terminal. The EPP is actually made up of multiple MEPPs, which arise when an activated terminal releases hundreds of neurotransmitter quanta. A series of EPPs, or a number of them stimulated simultaneously at many synapses, can then bring the cell to the threshold of the action potential. This combined action of EPPs is called summation.

      In contrast to electrical transmission, which takes place with almost no delay, chemical transmission exhibits synaptic delay. Recordings from squid synapses and neuromuscular junctions of the frog reveal a delay of 0.5 to 4.0 milliseconds between the onset of action potential at the nerve terminal and action potential at the postsynaptic site. This delay may be accounted for by three factors. First, diffusion of the neurotransmitter across the synaptic cleft takes approximately 0.05 millisecond. Second, the response of the postsynaptic receptor takes about 0.15 millisecond. This leaves 0.30 to 3.80 milliseconds for other processes. A third process, called mobilization of the transmitter, is traditionally postulated as taking up the remaining time, but evidence suggests that the time is occupied at least partially by the opening of calcium channels to allow the entry of Ca2+ into the presynaptic terminal.

      A series of nerve impulses arriving in rapid succession at the axon terminal is accurately reproduced as a series in the postsynaptic cell because the quanta of neurotransmitter released by each impulse are inactivated as soon as they stimulate the receptor proteins. Neurotransmitter inactivation is carried out by a combination of three processes. First, the neurotransmitter molecules simply diffuse out of the narrow synaptic cleft. Second, they are taken back into the presynaptic terminal by transmitter-sensitive transport molecules. Third, they are metabolized into inactive compounds by enzymes in the synaptic cleft.

Ion transport
      As is stated above, the lipid bilayer of the neuronal membrane tends to repel electrically charged, hydrated ions, making virtually impossible the movement across the membrane that is necessary for the generation of nerve impulses. The transmembrane movement of ions is actually carried out by molecular mechanism—specifically, by protein molecules embedded in the lipid layers. One mechanism, the sodium-potassium pump, maintains the resting potential, and another, the various ion channels, helps create the action potential.

      Since the plasma membrane of the neuron is highly permeable to K+ and slightly permeable to Na+, and since neither of these ions is in a state of equilibrium (Na+ being at higher concentration outside the cell than inside and K+ at higher concentration inside the cell), then a natural occurrence should be the diffusion of both ions down their electrochemical gradients—K+ out of the cell and Na+ into the cell. However, the concentrations of these ions are maintained at constant disequilibrium, indicating that there is a compensatory mechanism moving Na+ outward against its concentration gradient and K+ inward. This mechanism is the sodium-potassium pump. Actually a large protein molecule that traverses the plasma membrane of the neuron, the pump presents receptor areas to both the cytoplasm and the extracellular environment. That part of the molecule facing the cytoplasm has a high affinity for Na+ and a low affinity for K+, while that part facing the outside has a high affinity for K+ and a low affinity for Na+. Stimulated by the action of the ions on its receptors, the pump transports them in opposite directions against their concentration gradients.

      If equal amounts of Na+ and K+ were transported across the membrane by the pump, the net charge transfer would be zero; there would be no net flow of current and no effect on the membrane potential. In fact, in many neurons three sodium ions are transported for every potassium ion; sometimes the ratio is three sodium ions for every two potassium ions, and in a few neurons it is two sodium ions for one potassium ion. This inequality of ionic transfer produces a net efflux of positive charge, maintaining a polarized membrane with the inner surface slightly negative in relation to the outer surface. Because it creates this potential difference across the membrane, the sodium-potassium pump is said to be electrogenic.

      The sodium-potassium pump carries out a form of active transport—that is, its pumping of ions against their gradients requires the addition of energy from an outside source. That source is adenosine triphosphate (ATP), the principal energy-carrying molecule of the cell. ATP is formed by an inorganic phosphate molecule held in high-energy linkage with a molecule of adenosine diphosphate (ADP). When an enzyme in the pump, called sodium-potassium-ATPase, splits the phosphate from the ADP, the energy released powers the transport action of the pump.

Passive transport: membrane channels
      The sodium-potassium pump sets the membrane potential of the neuron by keeping the concentrations of Na+ and K+ at constant disequilibrium. The sudden shift from a resting to an active state, when the neuron generates a nerve impulse, is caused by a sudden movement of ions across the membrane—specifically, a flux of Na+ into the cell. Given the relative impermeability of the plasma membrane to Na+, this influx itself implies a sudden change in permeability. Beginning in the 19th century, researchers puzzled over the mechanism by which this change could occur. The idea arose that there must exist pores, or channels, through which the ions could diffuse, passing the barrier posed by the lipid bilayer. However, for years only the gross currents accompanying ionic movement could be measured, and it was only by inference that the presence of membrane channels could be postulated.

      The breakthrough came in the 1970s and '80s with the development of the patch-clamp technique, which enabled researchers to directly measure currents flowing across single ion channels in the membrane. The patch-clamp technique electrically isolates a small patch of neuron or muscle cell membrane by applying the tip of a micropipette filled with conducting solution to the membrane and forming a tight seal with it. As single channels in the patch undergo various transistional states between fully open and fully closed, the times of opening and closing are recorded and the amplitudes and duration of the currents are measured.

      Since the pioneering studies, the electrical and biochemical properties of certain channels have been characterized. Known as “voltage dependent” when activated by changes in the membrane potential and “neurotransmitter sensitive” when activated by neurotransmitter substances, these channels are protein structures that span the membrane from the extracellular space to the cytoplasm. They are thought to be cylindrical, with a hollow, water-filled pore wider than the ion passing through it except at one region called the selectivity filter. This filter makes each channel specific to one type of ion.

Sodium channels
      Voltage-sensitive sodium channels have been characterized with respect to their subunit structure and their amino acid sequences. The principal protein component is a glycoprotein containing 1,820 amino acids. Four similar transmembrane domains, of about 300 amino acids each, surround a central aqueous pore through which the ions pass. The selectivity filter is a constriction of the channel ringed by negatively charged carbonyl oxygens, which repel anions but attract cations. Also within the channel are thought to be two types of charged particles forming the gates that control the diffusion of Na+. One gate closes at polarization and opens at depolarization; the other closes at depolarization.

      It is thought that the resting, activated, and inactivated states of the sodium channel are due to voltage-dependent conformational changes in the glycoprotein component. These changes result from effects of the electrical field on the charges and dipoles of the amino acids within the protein. With a large electrical field applied to it, the protein has been observed to change its conformation from a stable, closed resting state to a stable, open state in which the net charge or the location of the charge on the protein is changed.

Potassium channels
      There are several types of voltage-dependent potassium channels, each having its own physiological and pharmacological properties. A single neuron may contain more than one type of potassium channel.

      The best-known flow of K+ is the outward current following depolarization of the membrane. This occurs through the delayed rectifier channel (IDR), which, activated by the influx of Na+, counteracts the effect of that cation by allowing the discharge of K+. By repolarizing the membrane in this way, the IDR channel restricts the duration of the nerve impulse and participates in the regulation of repetitive firing of the neuron.

      Another outward K+ current, occurring with little delay after depolarization, is the A current. IA channels are opened by depolarization following hyperpolarization. By increasing the interval between action potentials, they help a neuron to fire repetitively at low frequencies.

      Another type of potassium channel, the IK(Ca) channel, is activated by high concentrations of intracellular Ca2+. The opening of these channels results in hyperpolarization of the membrane, so that they appear to slow the repetitive firing of nerve impulses.

      The IM channel is opened by depolarization but is deactivated only by the neurotransmitter acetylcholine. This property may serve to regulate the sensitivity of neurons to synaptic input.

      A final type of potassium channel is the anomalous, or inward, rectifier channel (IIR). This channel closes with depolarization and opens with hyperpolarization. By allowing an unusual inward diffusion of K+, the IIR channel prolongs depolarization of the neuron and helps produce long-lasting nerve impulses.

Calcium channels
      As with potassium channels, there is more than one type of calcium channel. The inward calcium current is slower than the sodium current. There are at least two types of current in certain neurons of the central nervous system—a long-lasting current activated at positive potential and a transient current activated at more negative potential. There are two corresponding types of calcium channels: a large conductance channel that gives rise to a long-lasting current at positive membrane potentials and a low conductance channel that gives rise to a transient current at more negative potentials. In some neurons a third channel current occurs that is transient and can only be activated at high negative potential.

Anion channels
      There may be channels that pass anions such as Cl, but their existence is difficult to prove. Single-channel recordings of cultured tissue have shown selective Cl channels that are voltage dependent and of high conductance. Channels with lower conductance have been demonstrated in reconstituted artificial membranes as well as in neurons.

Neurotransmitters and neuromodulators
      The traditional models for the study of neurotransmitter release are either the neuromuscular junction of the frog, crayfish, and rat or the giant synapse of the squid. These synapses are relatively simple in their structure, with a single axon terminal forming an identifiable synapse at the postsynaptic membrane of a muscle fibre or neuron. Recordings can be obtained from these single-synaptic junctions in response to the release of a single neurotransmitter. At neurons of the central nervous system, on the other hand, the situation is more complex. Each central neuron has several synapses with other neurons at various locations, such as on the dendrites, soma, and initial segment of the axon. Several neurotransmitters, therefore—some excitatory and others inhibitory—may be involved in the final integrated response of a central neuron, making their identities difficult to determine.

      Further complicating neurotransmitter action is the presence not only of multiple transmitter substances but also of neuromodulators. Neuromodulators are substances that do not directly activate ion-channel receptors but that, acting together with neurotransmitters, enhance the excitatory or inhibitory responses of the receptors. It is often impossible to determine, in the presence of many substances, which are transmitters and which are modulators. Such is the case with many of the neuropeptides (see the section Neuroactive peptides (nervous system)).

      In addition to the multiplicity of transmitters and modulators there is a multiplicity of receptors. Some receptors directly open ion channels, while others activate the second-messenger system, any of a number of reactions that take place in the cytoplasm or plasma membrane and indirectly act upon the ion channels. One second-messenger system involves the activation by receptor proteins of linking proteins, which move across the membrane, bind to channel proteins, and open the channels. Another system is the cyclic adenosine monophosphate (cAMP) system. In this chain reaction, receptor proteins activate linking proteins, which then activate the enzymes that synthesize cAMP. The cAMP molecules activate other enzymes that, in turn, activate ion channels.

      Whether they activate channels directly or through a second-messenger system, neurotransmitters are considered to be primary messengers. Described below are the principal proved or suggested neurotransmitters of the mammalian nervous system and their corresponding receptors.

      Although early studies of acetylcholine were undertaken at neuromuscular junctions, where it is especially concentrated, the concept leading to the identification of the substance as a neurotransmitter of the central nervous system is a landmark in neuroscience. The concept is called Dale's principle after Sir Henry Dale (Dale, Sir Henry), a British physiologist who, in 1935, stated that a neurotransmitter released at one axon terminal of a neuron can be presumed to be released at other axon terminals of the same neuron. (Dale's principle refers only to the presynaptic neuron, as the responses of different postsynaptic receptors to a single neurotransmitter can vary in the same or different neurons.) The first application of Dale's principle was at the mammalian spinal cord, from which motor neurons send their axons to striated muscles, where the terminals are observed to release acetylcholine. According to Dale's principle, all the branches of a single motor neuron axon should release acetylcholine—including the terminals in the spinal cord. In fact, it was found that some collateral branches leave the motor axons and reenter the gray matter of the spinal cord, where they synapse onto spinal interneurons. The neurotransmitter released at these terminals is acetylcholine. High concentrations of the acetylcholine-synthesizing enzyme, choline acetyltransferase, and the enzyme for its breakdown, acetylcholinesterase, are also found in motor neuron regions of the spinal cord.

      Acetylcholine receptors (also called cholinergic receptors) appear in clusters on muscle-cell membranes opposite the active zones of presynaptic terminals. Their density at these receptor regions is between 7,000 and 30,000 sites per square micrometre (micron; millionth of a metre). The number drops drastically even a few nanometres (billionths of a metre) away from the receptor region, so that sensitivity to acetylcholine is about 50 to 100 times less one millimetre from the receptor region than it is at the receptor site itself. Cholinergic receptors also exist on the presynaptic terminals of neurons that release acetylcholine as well as on terminals that release other neurotransmitters. These receptors are called autoreceptors, and they probably regulate the release of neurotransmitter at the terminal.

      There are two main categories of cholinergic receptor, nicotinic and muscarinic. The nicotinic receptor is a channel protein that, upon binding by acetylcholine, opens to allow diffusion of cations. The muscarinic receptor, on the other hand, is a membrane protein; upon stimulation by neurotransmitter, it causes the opening of ion channels indirectly, through a second messenger. For this reason, the action of a muscarinic synapse is relatively slow. Muscarinic receptors predominate at higher levels of the central nervous system, while nicotinic receptors, which are much faster acting, are more prevalent at neurons of the spinal cord and at neuromuscular junctions in skeletal muscle.

 The nicotinic receptor channel is a glycoprotein composed of five subunits (see the figure—>). Two alpha- (α-) subunits contain the two acetylcholine-binding sites associated with the channel. Three other subunits—a beta- (β-) subunit, a gamma- (γ-) subunit, and a delta- (δ-) subunit—complete the protein. High-resolution electron microscopy with optical image reconstruction, as well as freeze-fracture electron microscopy, reveal a highly symmetrical structure, looking from the top somewhat like a life belt, with the presumed channel in the centre. About one-third of the protein protrudes from the plasma membrane, while the rest is embedded in the membrane or protruding into the cell.

      Patch-clamp techniques give information on single channel currents and, therefore, on the conductance and kinetics of the cholinergic receptor channel. At the neuromuscular junction, approximately 20,000 univalent ions carry the charge across a single activated channel, and a quantum of acetylcholine activates about 1,500 channels. The time constant for the decay of the MEPP is the same as that for channel closing. The time constant for channel closing is voltage dependent, with depolarization shortening the duration of open channels and hyperpolarization lengthening the duration.

      Studies show that nicotinic acetylcholine-activated channels allow cations to permeate the membrane with no specificity—that is, all cations can diffuse through the channels indiscriminately. Because the resting membrane is already near the equilibrium potential of K+, this means that much more Na+ and Ca2+ diffuse into the cell than K+ out, causing depolarization and excitation of the neuron or muscle cell. However, in certain molluscan neurons, nicotinic acetylcholine receptors can also activate Cl channels, causing hyperpolarization of the postsynaptic membrane and inhibition of excitability. With respect to muscarinic receptors, the situation is not clear. Second messengers may be involved, and potassium channels may be activated.

      These related hormones, also called adrenaline (epinephrine and norepinephrine) and noradrenaline, act to increase the heart rate, blood pressure, and levels of sugar and fat in the blood. They are secreted into the bloodstream by the adrenal glands (adrenal gland) in response to stress, but they are also synthesized and released as neurotransmitters by axon terminals in the central nervous system and in sympathetic fibres of the autonomic nervous system.

      Receptors sensitive to norepinephrine and epinephrine are called adrenergic receptors. They are divided into two types, α and β. These are further classified into subtypes α1, α2, β1, and β2.

      Both types of adrenergic receptors produce changes in the postsynaptic membrane potential by acting upon ion channels specific to K+ and Ca2+. They differ in the mechanisms that, upon stimulation by neurotransmitter, they employ to activate those channels. Stimulated β1 receptors bind to linking proteins that in turn bind to calcium channels, changing their shape and altering their permeability to the cation. More important, the linking proteins stimulate the synthesis of cAMP, which, through another series of reactions, opens potassium channels. The efflux of K+ tends to hyperpolarize the postsynaptic membrane, inhibiting the generation of a nerve impulse. The β2 receptor has been found on glial cells.

      The α2 receptor activates potassium channels in both the postsynaptic and presynaptic membranes, probably via linking proteins and the synthesis of cAMP. The α1 receptor acts on calcium channels through a series of reactions linked to the lipid molecules of the plasma membrane.

      Both epinephrine and norepinephrine are terminated by uptake back into the presynaptic terminals, where they are enzymatically degraded or inactivated.

       dopamine is a precursor of norepinephrine that acts as a neurotransmitter at certain synapses of the brain. Disorders at these synapses have been implicated in schizophrenia and Parkinson disease (parkinsonism).

      There are two types of dopaminergic receptors, called the D1 and the D2. The former catalyzes the synthesis of cAMP, and the latter inhibits its synthesis. These reactions then regulate calcium and potassium channels in the postsynaptic membrane. Dopaminergic receptors also exist on the presynaptic membrane. The neurotransmitter is terminated by uptake into the presynaptic terminal.

serotonin (5-hydroxytryptamine)
      Although the brain has only a small percentage of the serotonin found in the human body, there appears to be a strong relationship between the levels of this neurotransmitter at some regions of the brain and certain behavioral patterns, including sleep, sexual urge, and mood. At synapses of the peripheral nervous system, serotonin seems to prime muscle cells for an excitatory response to other neurotransmitters.

      Serotonin receptors, or 5HT receptors, activate calcium and potassium channels through linking proteins and the cAMP second-messenger systems. After acting on the postsynaptic receptors, the neurotransmitter is taken up by the presynaptic terminal and enzymatically degraded.

Amino acids (amino acid)
      Several amino acids (amino acid) exist in the central nervous system in extremely high concentrations, but their ubiquity makes their identification as true neurotransmitters difficult. Furthermore, because some of them are essential components of metabolic reactions, their presence within a neuron does not prove that they function as neurotransmitters. Nevertheless, there is enough evidence that some amino acids act as either excitatory or inhibitory transmitters. The excitatory amino acids include glutamic acid (or glutamate) and aspartic acid (or aspartate), and the inhibitory amino acids include gamma-aminobutyric acid (GABA) and glycine.

      Glutamate is the most abundant amino acid in the brain. Unlike acetylcholine, glutamate does not vary greatly in concentration from one region to the next. However, the dorsal gray matter of the spinal cord, which contains terminals of incoming dorsal roots, has large concentrations of glutamate. Aspartate, on the other hand, is believed to be concentrated in the interneurons of the ventral gray matter.

      At postsynaptic receptor sites glutamate depolarizes the membrane by opening nonspecific cation channels, which allow a net influx of Na+ and Ca2+. Of the excitatory amino acid receptors, the N-methyl-D-aspartic acid (NMDA) receptor has been thoroughly characterized. Patch-clamp studies show that this receptor is influenced by the presence of magnesium ions (Mg2+). In the absence of Mg2+, activated NMDA receptors open nonspecific cationic channels with no variation when the voltage is changed. With Mg2+ added to the extracellular medium, though, the frequency of channel openings is reduced when the membrane is hyperpolarized. Both glutamate and aspartate are probably inactivated by uptake systems at the presynaptic terminals and at glial cells surrounding some of the synaptic junctions.

      GABA and glycine cause hyperpolarization of the postsynaptic membrane. GABA is widely distributed in the brain, being especially prevalent at higher levels of the central nervous system. It is produced from glutamate by the enzyme glutamic acid decarboxylase (GAD). Consequently, the concentrations of GABA and GAD parallel each other in the nervous system.

      At postsynaptic receptor sites GABA opens chloride channels, causing in most cells a hyperpolarization of the membrane as Cl diffuses inward to reach its equilibrium potential. However, GABA inhibits presynaptic nerve fibres as well. At certain synaptic junctions the release of neurotransmitter is modulated by the binding to presynaptic receptors of neurotransmitter released from other neurons. An example of this is at the axon terminals of incoming dorsal roots in the dorsal gray matter. Projecting onto these terminals are other terminals that release GABA. Although GABA causes an increased Cl conductance at these terminals, the result is depolarization, not hyperpolarization, of the membrane. This is because the resting membrane potential of the receiving nerve terminal is much more negative than the Cl equilibrium potential. This means that as Cl flows into the terminal to reach equilibrium, the membrane is actually depolarized. The effect at the terminal is a decrease in neurotransmitter release.

      Unlike GABA, glycine is found mostly at lower levels of the central nervous system, including the spinal cord, medulla oblongata, and pons. It is a major inhibitor released by interneurons to suppress motoneuronal activity. Like GABA, glycine acts by increasing Cl conductance at the postsynaptic membrane, although it acts at a clearly different receptor.

      It appears that at least two molecules of glycine and GABA must bind to their respective receptors to activate a chloride channel. The action of both neurotransmitters is terminated by uptake back into the presynaptic terminal or into surrounding glial cells.

Neuroactive peptides
      Neuroactive peptides are sequences of amino acids, usually longer than amino acid neurotransmitters yet shorter than hormones or proteins. Unlike the classic neurotransmitters described above, which are formed by enzymes near the presynaptic terminals, neuroactive peptides are assembled by ribosomes attached to the endoplasmic reticulum. From there they are transferred to the Golgi apparatus, where they are packed into secretory vesicles and transported to the terminals. Some peptides are secreted by neuroendocrine cells of the hypothalamus or pituitary gland. Because they are released into the capillary system of the bloodstream and act at distant sites of the body, these are called neurohormones (neurohormone). Other peptides are released into the synaptic cleft between neurons of the central nervous system (including the hypothalamus). Many of these neuropeptides fulfill some criteria of neurotransmitters, evoking excitatory or inhibitory responses in postsynaptic ion channels, yet it is still uncertain to what extent they act as true neurotransmitters or as neuromodulators.

      Distinguishing neuropeptides from the classic neurotransmitters is the longevity of their action. While acetylcholine, for example, acts upon synaptic receptors for only a few milliseconds, neuropeptides have a course of action lasting from several seconds to several days. Also, neuropeptides are released in much lower concentrations than are transmitter substances, although the peptides have a much higher potency.

      The list of neuropeptides is not yet complete. Among those peptides known to affect synaptic transmission are substance P, neurotensin, somatostatin, vasoactive intestinal peptide, cholecystokinin, and the opioid peptides. The best-studied are the opioid peptides, so called because opiate (narcotic) drugs, such as morphine, are known to bind to their receptors and mimic their painkilling and mood-altering actions. All opoid peptides belong to three genetically distinct families: the β-endorphins, the enkephalins, and the dynorphins.

      It has long been known that opioids and opiate drugs have varied and powerful effects on pain, mood, sleep, sedation, and the cough reflex—apart from effects on the gastrointestinal tract and the cardiovascular system. It is not surprising, therefore, that there are multiple receptors for these substances. There may in fact be as many as eight different types of opioid receptors, but the four best-described are designated mu (μ), kappa (κ), delta (δ), and sigma (σ). The μ receptors, which readily bind morphine, are thought to mediate euphoria, respiratory depression, and physical addiction and to block pain pathways in the brain. The κ receptors bind preferentially to dynorphin and are thought to mediate analgesia and sedation at the spinal cord. The δ receptors, located primarily in the limbic portions of the brain, bind enkephalin. They may be responsible for dysphoria (extreme depression), hallucination, and respiratory and vasomotor stimulation. The σ receptors, found in the hippocampus, may be involved in alterations of affective behaviour, but their functions are unclear.

      The opioid receptors mediate their effects mainly by inhibiting regeneration of the nerve impulse at the postsynaptic membrane. They accomplish this by opening potassium channels or closing calcium channels, causing a net outflow of positive charge that keeps the postsynaptic membrane from reaching threshold potential. As with other neuropeptides, it is not known whether all the opioid receptors are activated by the opioids alone or by a combination of opioid and other transmitter substances. For this reason it is uncertain whether the opioid peptides are true neurotransmitters or are neuromodulators.

      The presence of peptides within certain structures of the central nervous system is well established; more important, peptides are often found in the same neurons with classic neurotransmitters or with other peptides. For example, substance P can be found in the same neurons of the brainstem as serotonin. In the sympathetic system, norepinephrine is found with somatostatin in some neurons and with enkephalin in others.

      Because some neuropeptides and neurotransmitters are stored in the same vesicles and secreted together in response to stimulation, a form of interaction between the substances appears likely. The interaction may take place at presynaptic terminals, altering the release of neurotransmitter, or it may take place postsynaptically, altering the effect of neurotransmitter. At the neuromuscular junction of the lobster, for example, the neurotransmitters serotonin and octopamine and the neuropeptide proctolin can act presynaptically to alter the amounts of GABA or glutamate released from the nerve terminals. In a similar manner, at some regions of the central nervous system opioid peptides inhibit the release of norepinephrine, acetylcholine, dopamine, and substance P.

      The discovery of more than one type of neuroactive substance in one set of axon terminals has disproved an assumption implied by Dale's principle—that a single neuron synthesizes and secretes a single neurotransmitter. Also called into doubt is another assumption—that a single neuron secretes a single set of neurotransmitters at all of its synapses. Researchers are finding evidence that different synapses of the same neuron act somewhat independently. This may mean that different areas of a single neuron synthesize different neuroactive substances. Such a phenonmenon would be another example of the metabolic and functional complexity of the nervous system.

Solomon D. Erulkar

evolution and development of the nervous system
      The study of the evolutionary development of the nervous system traditionally concentrated on the structural differences that exist at various levels of the phylogenetic scale, but certain functional characteristics, including biochemical and biophysical processes laid down early in evolution and amazingly well conserved to the present, can no longer be ignored. Two basic aspects of the evolution of the nervous system must be considered: first, how primitive systems serve newer functions, and, second, how the formation of new systems serves newer functional requirements.

      Early theories on the evolutionary origin of the nervous system argued for a three-stage process: first, the development of non-nervous “independent effectors,” such as muscle cells; second, the appearance of non-nervous receptors responding to certain modalities in a receptor-effector mechanism; and finally, the formation of a “protoneuron,” from which primitive nerves and ganglia evolved. This model is no longer considered valid. In primitive systems there appear to be many examples of non-nervous electrical conduction. For instance, large areas of epithelium covering the swimming bells in the hydrozoan order Siphonophora (which contains certain families of jellyfish) contain neither nerve nor muscle, yet depolarizing potentials between cells of the epithelium have been recorded. Similar examples from other systems in related orders suggest that the evolution of the nervous system may have begun with non-nervous epithelial tissue. The conduction of electrical potentials from one epithelial cell to the next may well have been via tight junctions, in which the plasma membranes of adjacent cells fuse to form cellular sheets. Tight junctions have low electrical resistance and high permeability to molecules. They also occur in large numbers in embryos, suggesting that the electrical potentials of cells joined in this manner serve as a driving force for the movement of ions and even nutritive substances from one cell to the next. These phenomena suggest that electrically mediated junctional transmission is older than chemically mediated synaptic transmission, which would require that some epithelial cells secrete chemical substances.

      Many investigators believe that neurons originated from endothelial secretory cells that could secrete chemical substances, respond to stimulation, and conduct impulses. Specialization may then have brought about an outer receptor surface and an inner conducting fibre. In fact, neurosecretory cells (neurosecretory cell) can propagate action potentials (action potential), and many neurons secrete chemical substances, called neurohormones, that influence the growth and regeneration of cells at other sites of the body. Some researchers suggest that neurons may have first appeared as neurosecretory growth-regulating cells in which elongated processes were later adapted to rapid conduction and chemical transmission by release of neurotransmitters at their endings.

      There is an amazing consistency in neurotransmitters (neurotransmitter) present in different organisms of a given phylum, although different phyla may show striking differences. Thus, in vertebrates (vertebrate), including fishes, amphibians, reptiles, birds, and mammals, the motor neurons (neurons whose fibres innervate striated muscle) are always cholinergic (that is, they secrete the neurotransmitter acetylcholine). In arthropods (arthropod), on the other hand, they are not cholinergic, although the sensory neurons do secrete acetylcholine. The number of known neurotransmitters in the animal kingdom is small, and their presence in more primitive organisms as well as in nervous systems of later vertebrates shows a striking conservation of these substances throughout evolution.

      If later organisms evolved from single-celled ancestors, then there must have been some system for the transmission of information from one evolutionary stage to the next. These conditions have been defined as: (1) a stable means for encoding, transmitting, and decoding characteristics from one generation to the next, (2) the possibility of alterations in the code taking place by mutation or sexual recombination, and (3) a means of selecting (natural selection) only those characteristics for transmission that are favourable for survival. As mentioned in the section Stimulus-response coordination (nervous system), protozoans (single-celled organisms) move toward places that are favourable for survival, such as areas with optimal conditions of light and temperature. As the metazoans (multicelled organisms) developed, entire groups of cells probably tended to move toward favourable conditions, and when the number of cells became very large, a system of internal communication—in effect, a nervous system—developed. Two general types developed: the diffuse nervous system and the centralized nervous system.

Diffuse nervous systems
      The diffuse nervous system is the most primitive nervous system. In diffuse systems nerve cells are distributed throughout the organism, usually beneath the outer epidermal layer. Large concentrations of nerve cells—as in the brain—are not found in these systems, though there may be ganglia (ganglion), or small local concentrations of neurons. Diffuse systems are found in cnidarians (cnidarian) (hydroids (hydroid), jellyfish, sea anemones (sea anemone), corals (coral)) and in ctenophores (ctenophore), or comb jellies. However, the primitive nervous systems of these organisms do not preclude prolonged and coordinated responses and integrated behaviour to the simplest stimuli. An example is the movement of the sea anemone Calliactis onto the shell of the hermit crab Pagurus in response to a factor present in the outer layer of the empty mollusk shell occupied by the crab. This movement requires integration of the highest order.

 Most cnidarians, such as those of the genus Hydra, have what is called a nerve net—a meshlike system of individual and separate nerve cells and fibres dispersed over the organism (see the diagram—>). Species of Hydra have two nets, one located between the epidermis and the musculature and the second associated with the gastrodermis. Connections occur at various points between the two nets, with individual neurons making contact but not fusing, thereby forming structures similar to the chemically mediated synapses of vertebrates. Several specializations occur within various species. In Hydra the neurons are slightly more concentrated in a ring near the pedal disk and the hypostome (the “mouth”), but in jellyfish of a related genus the nerve fibres form a thick ring at the margin of the bell to form “through” conduction pathways.

      The nervous systems of cnidarians correspond to their radially symmetrical bodies, in which similar parts are arranged symmetrically around a hollow gut cavity called the coelenteron. In some species nerve fibres course along the radial canals, where there may be arranged sensory bodies, called rhopalia, which contain ganglionic concentrations of neurons. In the sea anemone Metridium some of the nerve fibres are seven to eight millimetres (three inches) long and form a system for fast conduction of nerve impulses. Such specializations may have allowed the evolution of different functions. Rapid coordination of swimming movements requires a fast-conducting pathway, while feeding relies on the nerve net. Integrative activity is likely to occur at the sensory ganglia, which may represent the first forms of a centralized nervous system.

      The terminals forming synapselike structures in nerve nets contain synaptic vesicles that are believed to be packed with neurotransmitters and neuroactive peptides. Peptides present in Hydra nervous systems also exist in mammalian systems as neuromodulators, neurohormones, or even possible neurotransmitters.

      Transmission in the nerve net is relatively slow compared with that in other nervous systems (0.04 metre per second in radial fibres of Calliactis compared with 100 metres per second in some fibres of the dog). Many repetitive stimuli may be required to elicit responses at these synapses. Long refractory periods are also characteristic of nerve nets, having durations about 150 to 300 times those seen at mammalian nerve fibres.

      Finally, pacemaker systems are present in animals with nerve nets. In the sea anemone Metridium these systems are expressed in a series of spontaneous rhythmic movements that occur in the absence of any detectable stimulus. It is not known whether the movements originate from a “command” neuron or group of neurons or whether they arise without neuronal stimulation. It has been postulated that pacemaking cells were present in epithelial conducting systems known not to be nervous but that eventually evolved into neuronal tissue.

Centralized nervous systems
      The development of the nerve net allowed an organism to engage in several different behaviours, including feeding and swimming. The development in the net of rapidly conducting bundles of fibres and of pacemaker systems allowed rapid withdrawal and rhythmic swimming activities, respectively, in some cnidarians. However, it is at the level of the flatworms (flatworm) (phylum Platyhelminthes) that there appears a longitudinal nerve cord and an anterior collection of nerve cells that can be called a brain. Furthermore, there are well-defined sensory and motor pathways as well as coordinating interneurons. Although nerve nets and pacemaker activity are still present in the flatworms, the presence of ganglia or a brain concentrated at the cephalic (head) end of the organisms represents a simple beginning to the complex centralized systems that develop at higher levels of the phylogenetic tree.

Simple bilateral systems
 The flatworms were the first invertebrates (invertebrate) to exhibit bilateral symmetry and also the first to develop a central nervous system with a brain. The nervous system of a free-living flatworm such as Planaria (see the diagram—>) consists of a brain, longitudinal nerve cords, and peripheral nerve plexuses (interlacing networks of peripheral nerves; from Latin plectere, “to braid”). Located in the anterior portion of the animal, the brain is composed of two cephalic ganglia joined by a broad connection called a commissure. Longitudinal nerve cords, usually three to five pairs, extend posteriorly from the brain; they are connected by transverse commissures, and smaller, lateral nerves extend from the cords. The lateral nerves give rise to the peripheral nerve plexuses. The submuscular nerve plexus, consisting of sensory cells, ganglion cells, and their processes, is situated in the loose tissue (mesenchyme) below the subepidermal musculature. Another subepidermal plexus is located at the bases of the epithelial cells above the muscular layer.

      Planaria (planarian) are richly supplied with sensory receptors. Single sensory cells in the nerve plexuses are widely scattered over the organism. Sensory organs also are present and include ciliated pits and grooves, auricles, the frontal organ, statocyst, and eyes. The ciliated pits and grooves contain chemical receptors, or chemoreceptors, which permit the animal to detect food. The statocyst is responsible for balance and such reactions as rising to the surface of the water or sinking. The eyes, or ocelli, may occur as a pair situated anteriorly or may be scattered abundantly over the head region depending on the species. Short optic nerves connect the eyes with the brain.

      Seven types of nerve cell bodies and two types of neuroglia have been described in Planaria. Removal of the brain results in the abolition of such functions as food finding and recognition and severe deficits in locomotion. However, the nerve cords by themselves can mediate a certain amount of locomotion as well as righting and avoidance reactions.

      Nematodes (nematode) (phylum Aschelminthes) have a high degree of centralization, with three-quarters of all nerve cells concentrated in a group of anteriorly placed ganglia and no peripheral plexuses or nets. They usually have eight longitudinal cords, commissures between dorsal and ventral cords, six cephalic nerves, a few special ganglia and nerves in the tail, and two sympathetic systems (one anterior and one posterior).

Moderately cephalized (cephalization) systems
      Basic similarities in the nervous systems of the annelid worms, mollusks, and arthropods include an anteriorly situated brain, connectives running from the brain around the esophagus and joining paired longitudinal cords, and ventral nerve cords with ganglia along their length. The trend toward greater centralization and cephalization of nervous functions is continued within these groups, reaching its peak in the higher mollusks and arthropods.

Annelids (annelid)
 The brain of most annelids (annelid) (phylum Annelida; segmented worms, including the leeches (leech) and terrestrial earthworms (earthworm)) is relatively simple in structure. The earthworm brain is a bilobed mass lying above the pharynx in the third body segment (see the diagram—>). Sensory nerves leave the brain and run forward into the prostomium (extreme anterior end) and first segment. The brain of the active, predatory polychaetes (polychaete) (a class of marine worms) is more complicated. In some, the brain can be divided into a forebrain, midbrain, and hindbrain; a single pair of circumesophageal, or circumpharyngeal, connectives leave the brain, surround the anterior gut, and connect with the ventral nerve cord.

      The most primitive annelids have a pair of ventral nerve cords joined by transverse connectives; the most advanced forms have the cords fused to form a single cord. A ganglionic swelling of the cord is found in each body segment, with the most anterior ganglion, the subpharyngeal ganglion, being the most prominent. Two to five pairs of lateral nerves leave each ganglion to innervate the body wall of that segment. A subepidermal nerve plexus occurs over the whole body. Another plexus, called the enteric (digestive nerve plexus), stomodaeal, or sympathetic system, is found in the wall of the gut.

      Giant axons, usually few in number, travel the length of the cord. They may belong to one cell or be composed of many neurons. These axons are capable of very rapid conduction of impulses to the segmental muscles; their main function is to permit the worm to contract very rapidly as a defense against predators.

      The usual slow crawling movements of worms are mediated by a series of reflex arcs. During crawling, the contraction of muscles in one segment stimulates stretch receptors in the muscle. Impulses are carried over sensory nerves to the cord, causing motor neurons to send impulses to the longitudinal muscles, which then contract. The longitudinal pull activates stretch receptors in the following segment, and a wave of contraction moves along the worm.

      Studies of the nervous systems of annelids show certain behavioral capabilities, including perception, motor coordination, and learning. Because the neuronal organization behind these capabilities can be deduced, they may give an indication of the mechanisms underlying similar patterns of activity and behaviours at other levels of the phylogenetic scale.

      Two rhythmic movements generated by the leech, the heartbeat and swimming rhythm, have been extensively studied. The coordinated heartbeat rhythm is produced by heart excitor motor neurons, which show rhythmic activity in which bursts of action potentials alternate with bursts of inhibitory synaptic potentials derived from rhythmically firing inhibitory interneurons. The heartbeat appears to be produced by a central rhythm generator. The swimming movement, on the other hand, is generated by a neuronal network requiring many more cells. These neuronal oscillators may form the basis for neuronal generators of rhythmic movements in other animals at higher levels of the phylogenetic scale.

Simple mollusks (mollusk)
      The nervous systems of the more primitive mollusks (mollusk) (snails (snail), slugs (slug), and bivalves (bivalve), such as clams (clam) and mussels (mussel)) conform to the basic annelid plan but are modified to conform with the unusual anatomy of these animals. In snails a pair of cerebral ganglia constitutes the brain, which overlies the esophagus. Nerves leave the brain anteriorly to supply the eyes, tentacles, and a pair of buccal ganglia. These last ganglia, also called the stomatogastric head ganglia, innervate the pharynx, salivary glands, and a plexus on the esophagus and stomach. Other nerve cords—the pedal cords—leave the cerebral ganglia ventrally and terminate in a pair of pedal ganglia, which innervate the foot muscles. Another pair of nerve cords—the visceral cords—leave the brain and run posteriorly to the visceral ganglia. The pleural ganglion, supplying the mantle, or fleshy lining of the shell, and the parietal ganglion, innervating the lateral body wall and mantle, are located along the visceral nerves. Intestinal ganglia connected with the pleural ganglia innervate the gills, osphradium (a chemical sense organ), and mantle. Sense organs of snails include eyes, tentacles, statocysts, and osphradia.

      In the bivalves (bivalve), a cerebropleural ganglion is situated on either side of the esophagus. An upper pair of nerve cords leaves these ganglia and runs posteriorly to the visceroparietal, or visceral, ganglia. The visceral ganglia supply the mantle, adductor muscles (which close the shell), and internal organs. A second pair of nerve cords travels ventrally to the pedal ganglia. Most of the sense organs are found at the edge of the mantle. In the scallop, for example, the eyes (eye, human) are set in a row. They are well developed and consist of a cornea, a lens, and a retina, in which the photoreceptor cells are not placed superficially (an arrangement much like that in the vertebrate retina).

      Elementary forms of learning and memory have been studied at a cellular level by analysis of the neuronal activity of the marine snail (Aplysia californica). This simple mollusk withdraws its gill and siphon in response to a mild tactile stimulus. The neural circuit for this reflex consists of a sensory component from the siphon that forms single-synapse junctions with motor neurons that cause the gill to withdraw. The sensory cells also project onto interneurons whose outputs converge onto the same motor neurons. In response to a stimulus, the sensory neurons generate large excitatory postsynaptic potentials at both interneurons and motor neurons, causing the generation of action potentials in the motor neurons that in turn cause the gill to withdraw. When the stimulus is repeated many times, the postsynaptic potentials become reduced in size and the response becomes weaker. Finally, the postsynaptic potentials become so small that action potentials are no longer generated and the gill no longer responds. This reduced behavioral response is known as habituation. Habituation may be caused by the closing of calcium channels, which decreases calcium influx into the presynaptic terminals and, therefore, decreases neurotransmitter release. Other evidence suggests that habituation results from fewer neurons in the network being activated.

      Another behavioral paradigm, sensitization, has also been examined in Aplysia. In sensitization the reflex activity increases in strength with added stimulation. The mechanism underlying this response is presynaptic facilitation, which is thought to be caused by an increase in the second messenger cAMP in the terminals of the sensory neurons.

      These two examples—habituation and sensitization—show that important features of a more complex nervous systems can be studied in organisms at lower stages of evolution. First is what can be called the plasticity (neuroplasticity) of the nervous system, the phenomenon of changes occurring in the strength of synaptic responses. Changes in synaptic efficacy may underlie certain mechanisms for short- and long-term memory—even in more complex animals such as humans. Changes in the structure of the synapse may be a long-term effect of plasticity. For example, the numbers of active zones at nerve terminals are reduced with long-term habituation but increased with long-term sensitization. Finally, the molecular mechanisms underlying these changes may be the same or at least similar at all levels of the phylogenetic tree. Habituation of the escape response has been seen in polychaete worms, cockroaches (cockroach), and crayfish.

Complexly compartmentalized systems
      The highest degree of development of the invertebrate nervous system is attained by the cephalopods (cephalopod) (squids (squid), cuttlefishes (cuttlefish), and octopuses (octopus)) among the mollusks (mollusk) and by the insects (insect) and spiders (spider) among the arthropods (arthropod). Although the basic plan of these nervous systems is similar to that of the annelids, there are several advances. First, there is a high degree of cephalization, with nervous functions concentrated in the head region of the animal. In addition, ganglia are fused and farther forward, and nerve cells, less abundant in the peripheral nervous system, are situated in the brain or ganglia so that the nerve cords consist only of nerve fibres. Finally, control and coordination of specific functions, such as locomotion and feeding, are compartmentalized in particular parts of the nervous system.

Complex mollusks
      The complex nervous system of the cephalopods is correlated with the active movement and predatory habits of these organisms. Most of the ganglia typical of mollusks are concentrated or fused in a brain that encircles the esophagus. Nerves extend from the brain to ganglia at the base of the arms or tentacles and from the ganglia the length of the arms. A pair of large pallial nerves connects the brain with a pair of stellate ganglia on the inner surface of the mantle. The stomatogastric ganglia supply nerves to the digestive tract.

      A great variety of functions are centralized in the brain and compartmentalized to specific brain regions. These activities may be local, simple, and uncoordinated with other regions or may be extensive, complex, and coordinated, involving large groups of muscles. The highest centres of the cephalopod brain are the associative areas, which are thought to be involved with discrimination between objects, learning, and memory.

      The giant-fibre system—also seen in earthworms and insects—is very well developed in the squid. The diameter of giant fibres is many times greater than the diameter of most other nerve fibres. Giant neurons in the brain send fibres to the retractor muscles of the head and the funnel or to the stellate ganglion. Fibres from the stellate ganglion fuse to form giant fibres that innervate the mantle. Because of their large size, these fibres are capable of rapid conduction, which, in turn, permits extremely rapid movement.

      The eyes (eye, human) of cephalopods are especially well developed and bear close resemblance to the vertebrate eye. The eye fits into a socket of cartilaginous plates separate from the cartilages that protect the brain, and external muscles permit its movement. A transparent cornea covers the surface and can be focused for both near and far objects. There is a pupil formed by an iris diaphragm, which can regulate the amount of light reaching the retina. The retina contains light-sensitive cells. The axons of the photoreceptors, or rod cells, form the optic nerves, which terminate in the extremely large optic lobes of the brain.

      The cephalopods are strikingly different in many respects from other molluscan classes. The nervous system as described above is more highly developed and, consequently, the behavioral repertoire much more complex. First, the animals are predators; they move, they use their eyes in search of food, they use receptors in their arms for detection of tactile or chemical stimuli, and they have exceptionally fast muscle action. Second, they have an enormous flexibility of response, discriminating between palatable and unpalatable prey and “learning” to attack or not to attack. They can also change colour to blend into their environment if needed.

      The mollusks as a whole provide an important link in the developing complexity of the nervous system. Indeed, the presence in their systems of vertebrate as well as natural molluscan neuroactive peptides may give some clue to the true place of these animals in the phylogenetic scale.

Arthropods (arthropod)
 The other complex compartmentalized nervous system is found in arthropods (see the diagram—>). The arthropodan brain consists of three main regions: the protocerebrum, deutocerebrum, and tritocerebrum. The anterior protocerebrum, which receives the nerves of the eyes and other organs, contains centres, or neuropils, such as the optic centres and bodies known as corpora pedunculata. The neuropils function as integrative systems for the anterior sense organs, especially the eyes, and in control of movement; they also are the centres for the initiation of complex behaviour. The deutocerebrum contains the association centres for the first antennae. The posterior tritocerebrum contains association neuropils for the second antennae (of crustaceans) and gives rise to nerves that innervate the mouthparts and the anterior digestive canal. The latter constitute the stomatogastric system, which regulates the intake of food and the movement of the gut necessary for digestion. This system bears a resemblance to the vertebrate autonomic system.

      The ventral nerve cord, connected to the brain by the circumesophageal connectives, is composed of a double row of ganglia connected longitudinally by connectives and transversely by commissures. Different groups of arthropods exhibit different degrees of fusion of the ganglia. In insects the first ganglion, the subesophageal, is formed by fusion of three pairs of ganglia; it sends nerves to the mouthparts and to the salivary glands. The segmental ganglia in the thorax and abdomen provide nerves to the appendages, dorsal muscles, sense organs, and heart. Insects have 3 pairs of thoracic ganglia and up to 10 abdominal ganglia.

      The most common sensory receptors in arthropods are the cuticular hairs, many of which are mechanoreceptors, sensitive to touch, vibration, water currents, or sound waves; some hairs are chemoreceptors, which detect odours or chemicals in the water. Hairs situated near the joints are stimulated by body movements and thus provide a sense of the position of the joint or appendage during locomotion or flight. Many sensory cells and organs are concentrated in the antennae, and a statocyst is found at the base of each antenna.

      Spiders have several pairs of simple eyes with cup-shaped retinas. Crustaceans (crustacean) and insects, however, have a pair of well-developed compound eyes, each consisting of a large number of visual units called ommatidia. Each ommatidium contains six to eight sensory receptors arranged under a cornea and refractile cone and is surrounded by pigment cells, which adjust the intensity of light. Each ommatidium can act as a separate eye and is capable of responding to its own visual field. Such an arrangement seems particularly well suited for detecting movement across a wide visual field.

      In spite of the small size of insects, some of their nerve cells and axons are larger in diameter than any neuron in the human nervous system. The number of insect neurons is relatively small, so that each neuron must be capable of dealing with a maximum amount of information. Evasive behaviour requires mechanisms of this sort—as exemplified by the jumping muscle of the grasshopper, which is supplied by only a few motor axons, while those muscles of a human required for a similar purpose have tens of thousands of axons. Another example is the ear of a noctuid moth (owlet moth). Each ear is essentially a tympanic membrane forming the outer wall of an air-filled cavity in the thorax. A five-tissue strand, the acoustic sensillum, runs from the centre of the tympanic membrane across the tympanic cavity to a nearby skeletal support. This sensillum has two acoustic sensory receptors, called A cells. From the central end of each A cell, an axon passes within the sensillum to the skeletal support and then in the tympanic nerve to the thoracic ganglia of the moth.

      The A cells encode the intensity of ultrasound by the frequency with which they fire action potentials. Each ear is capable of responding differently to different stimuli, so that differences between the two ears in the duration of the action potentials and response times could allow for binaural detection of the source of a sound. The impulses are conducted to the pterothoracic ganglion, where they must influence muscles used for avoidance of predators. The important point here is that two A cells provide enough information to allow the moth to take evasive behaviour.

      Clearly, simplicity of the neural circuitry is required for speedy response; while some information may be sacrificed, escape will not be compromised. In fact, the quick evasion of predators has probably influenced the evolution of the giant-fibre systems of worms and squid as well as crustaceans and insects. These giant fibres conduct impulses at much higher conduction velocities than do smaller axons, while the information-handling capacity of many small axons acting together is far greater than that of giant axons. Under these circumstances it is clear that different systems evolved in the invertebrate nervous system in response to different qualities of stimuli in the environment to which the organism had to react—one responsible for survival and the other for information.

      Neurosecretory cells (neurosecretory cell), which have been identified in all the major invertebrate groups, reach their highest degree of development in the arthropods. The principal system of insects consists of neurosecretory cells in the protocerebrum of the brain. The axons of these cells form nerves that innervate structures called corpora cardiaca, situated just posterior to the brain. A corpus cardiacum is an organ in which neurosecretory products are stored in the nerve terminals for later release into the vascular system.

      The neurosecretory cells of the protocerebrum manufacture a hormone called ecdysiotropin (thoracotropic hormone), or brain hormone. Sensory nerve impulses reaching the brain regulate the release of the hormone from the nerve endings into the blood within the corpora cardiaca. The hormone then stimulates a non-neural endocrine gland, the ecdysial gland, located in the thorax. The ecdysial glands in turn release the hormone ecdysone, which initiates molting during larval development and also stimulates differentiation into adult tissues. Another hormone, however, the juvenile hormone, keeps tissue in a juvenile or larval form. This hormone is released by the corpora allata, another pair of non-neural endocrine glands, located behind the corpora cardiaca. The successive life stages in insect development are, therefore, determined by the varying levels of ecdysone and juvenile hormone.

The vertebrate system
      The nervous system of vertebrates (vertebrate) has two main divisions: the central nervous system, consisting of the brain and spinal cord, and the peripheral nervous system, which in humans includes 12 pairs of cranial nerves (cranial nerve), 31 pairs of spinal nerves (spinal nerve), and the autonomic (autonomic nervous system), or involuntary, nervous system.

      Anatomic structures such as the nervous system are described according to their postion. In four-legged animals the upper (back) surface is called dorsal and the lower (belly) surface ventral. The terms anterior, cranial, cephalic, and rostral refer to the head end of the body, posterior and caudal to the tail end. In humans, since they stand erect, the situation is more complicated: dorsal becomes equivalent to posterior, and ventral is the same as anterior; cranial is often called superior, and caudal inferior. Objects near the middle plane of the body are medial and those farther away are lateral. Proximal refers to structures nearest the central bulk of a structure and distal to ones away from it. In referring to another structure, if it is located on the same side of the body, it is known as ipsilateral; if it is on the opposite side, it is contralateral.

      Neurons are often gathered into localized masses. In the peripheral nervous system these accumulations are called ganglia; in the central nervous system they are called nuclei. Portions of the central nervous system in which unmyelinated neurons and neuroglia predominate are called gray matter; areas in which myelinated neurons dominate are called white matter. Efferent, or motor, nerve fibres carry impulses away from the central nervous system; afferent, or sensory, fibres carry impulses toward the central nervous system. Visceral fibres innervate the viscera such as the heart and intestines, and somatic fibres innervate the body-wall structures such as skin and muscle. In the central nervous system the nerve fibres are organized in bundles called tracts, or fasciculi. Ascending tracts carry impulses along the spinal cord toward the brain, and descending tracts carry them from the brain or higher regions in the spinal cord to lower regions. The tracts are often named according to their origin and termination; for example, the corticospinal tract consists of fibres running from the cerebral cortex in the brain to the spinal cord.

The primitive condition
      The vertebrates constitute an advanced subdivision of the phylum Chordata. All chordates (chordate) at some time in their life have a rodlike bar called the notochord running the length of the body. Lower chordates (acorn worms, tunicates (tunicate), and amphioxus), which lack a vertebral column, illustrate the most primitive features of the chordate nervous system. In these animals the nerve cord is a rather uniform-appearing dorsally placed tube with a hollow cavity, which corresponds roughly to the spinal cord of the vertebrates, suggesting that the spinal cord is the most primitive component of the central nervous system.

      In amphioxus and in lower vertebrates such as lampreys (lamprey), the sensory fibres and motor fibres leave the cord in dorsal and ventral roots to supply the adjacent body segments called myotomes. The dorsal and ventral roots remain separate nerves and arise at alternate positions along the cord. In lower fishes there is still alternation of dorsal and ventral roots, but the roots unite in a single spinal nerve. In higher vertebrates the two roots unite in a single spinal nerve and leave the cord at the same level, one above the other. Each spinal nerve supplies a single myotome. When appendages (fins, wings, arms, and legs) develop from several myotomes, the nerves continue to supply their original segments, and branches of the spinal nerves become interwoven to form plexuses.

      The brain of vertebrates developed by the accumulation of nerve cells at the cephalic end of the nerve cord. At first this diffuse collection of nerve cells regulated the reflex activity of spinal motor neurons. These cells are comparable to the reticular formation occupying the brainstem of higher vertebrates. The brainstem, thus, is the oldest portion of the brain.

Encephalization
      Early in the evolution of vertebrates, a special sensory system became associated with each major part of the brain: the olfactory organs with the forebrain, the eye with the midbrain, and the ear and related organs with the hindbrain. Each of the three sections, furthermore, developed dorsal outgrowths of gray matter forming, respectively, the cerebrum, the midbrain roof, or tectum, and the cerebellum. With these developments the three-part brainstem was then transformed into a brain of five regions: telencephalon, diencephalon, mesencephalon, metencephalon, and myelencephalon. The addition of these nerve centres to the primitive brain stem allowed greater coordination and association between the sensory and motor fibres.

 In tracing the development of the parts of the brain in the different vertebrate classes, some general features are apparent. There is a correlation between the size of a particular part of the brain and its importance in the functions of an animal. Some neural structures (e.g., the olfactory bulb; see the diagram—>) have considerable size and importance in more primitive animals but are less conspicuous in most recent animals. Progressing from primitive to recent animals, there is a gradual cephalic shift of function from the lower brain stem to the higher cerebral cortex.

      The hindbrain is comparable to an enlarged, anterior section of the spinal cord. In the gray matter are dorsal sensory and ventral motor columns similar to those present in the cord. The longitudinal continuity of these columns is preserved in the earlier vertebrates, but in more recent vertebrates the columns break up into discrete nuclei that serve some of the cranial nerves. The hindbrain exerts partial control over the spinal motor neurons through the reticular formation. Fish and tailed amphibians, in addition, have a pair of giant cells called the cells of Mauthner, which exert some control over the local spinal-cord reflexes responsible for the rhythmic swimming undulations and the flip-tail escape response characteristic of these animals.

      The hindbrain is the area of reception of one of the main sensory systems, the acoustico-lateralis system, which consists of the ear (ear, human) (hearing and equilibrium) and the lateral-line organs (vibration and pressure). The latter, situated in rows along the head and body, are retained in fish but disappear in the land vertebrates.

 The cerebellum originated as a specialized part of the acoustico-lateralis area. The oldest part of the cerebellum—the archicerebellum—is concerned with equilibrium and connected with the inner ear and the lateral-line system. The anterior lobe of the cerebellum represents the paleocerebellum, an area that regulates equilibrium and muscle tone; it constitutes the main mass of the cerebellum in fish (see the diagram—>), reptiles, and birds. In mammals the development of the cerebral cortex and its connections with the cerebellum are correlated with the appearance of the large cerebellar hemispheres. This new part of the cerebellum, or neocerebellum, coordinates skilled movements initiated at cortical levels. In mammals a great mass of fibres connects the brain stem to the cerebellum; this region forms the pons, which, together with the cerebellum, constitutes the metencephalon. The caudal part of the hindbrain remains as the medulla oblongata (myelencephalon).

  The midbrain (mesencephalon) and diencephalon constitute the anterior portion of the brain stem. Sensory and motor nuclei for cranial nerves extend from the hindbrain to the midbrain. The roof of the midbrain, or tectum, developed as the primary visual centre. The optic lobes, especially prominent in fish and birds (see the diagram—>), are a part of this area. In fish and amphibians (see the diagram—>) the tectum is the major centre of the nervous system and wields the greatest influence on body activity. While this area is still significant in reptiles and birds, it is supplanted in importance by the cerebral hemispheres. In mammals most of the optic sensations are relayed to the cerebral cortex. With development of the cerebral cortex, the thalamus becomes less significant as an association area and more important as a relay centre for sensory impulses. Centres for visceral sensations and visceral motor responses become established in the hypothalamus.

Dominance of the cerebrum
  Ascending the vertebrate scale, the cerebral hemispheres become more and more important as association centres. The cerebral hemispheres begin their development as paired outgrowths of the forebrain and serve as centres of olfactory reception. In the older vertebrates the forebrain is divided into the olfactory bulb—where the olfactory nerve fibres end—and the cerebral hemisphere. The hemisphere at this time, referred to as the paleopallium, is merely an olfactory lobe serving as an association area for olfactory impulses. The olfactory lobes are prominent in animals such as amphibians, but in birds and primates in which sense of smell is less important the lobes are reduced in extent. In amphibians the hemispheres consist of three parts: the paleopallium (olfactory lobe), archipallium, and basal nuclei. All three areas receive olfactory stimuli and discharge impulses to the brain stem. The archipallium is a correlation centre and forerunner of the mammalian hippocampus. The basal nuclei are equivalent to the corpus striatum and function as association areas with connections with the thalamus. In primitive reptiles the basal nuclei have moved to the inner part of the hemisphere, whereas the other areas of gray matter have moved toward the surface. In advanced reptiles a new association centre, the neopallium, appears between paleopallium and archipallium. In birds there is nothing corresponding to the neopallium, but the basal ganglia (that is, the corpus striatum) are enormously expanded. In mammals the neopallium becomes greatly enlarged to exceed all the other parts of the brain in size. This region assumes more and more of the higher types of neural activity in correlation, association, and learning. At first the neopallium expands to envelop the other brain structures. The archipallium becomes folded into a small area on the median part of the hemispheres, where it remains as the hippocampus. The paleopallium (olfactory lobes) constitutes a small ventral region on the hemisphere (see the diagram—>), the pyriform lobe. The corpus striatum (old basal nuclei) becomes a central part of the hemisphere. Further expansion of the neopallium in primates and humans (see the diagram—>) causes extensive folding and results in a very convoluted surface of the brain.

Thomas L. Lentz Solomon D. Erulkar

Additional Reading
General physiological features of the nervous system are presented in John G. Nicholls, A. Robert Martin, and Bruce G. Wallace, From Neuron to Brain: A Cellular and Molecular Approach to the Function of the Nervous System, 3rd ed. (1992); H.H. Dukes, Dukes' Physiology of Domestic Animals, 11th ed., edited by Melvin J. Swenson and William O. Reece (1993); Eric R. Kandel, James H. Schwartz, and Thomas M. Jessell (eds.), Principles of Neural Science, 4th ed. (2000); Arthur C. Guyton and John E. Hall, Textbook of Medical Physiology, 10th ed. (2000); Gordon M. Shepherd, Neurobiology, 3rd ed. (1994); and Alan Peters, Sanford L. Palay, and Henry DeF. Webster, The Fine Structure of the Nervous System: Neurons and Their Supporting Cells, 3rd ed. (1991), a description of nerve cells based on numerous electron micrographs.The transmission of information in the nervous system via electrical signals is analyzed in three works of historical interest: Keith Lucas, The Conduction of the Nervous Impulse, rev. by E.D. Adrian (1917); E.D. Adrian, The Mechanism of Nervous Action: Electrical Studies of the Neurone (1932, reissued 1959); and Joseph Erlanger and Herbert S. Gasser, Electrical Signs of Nervous Activity, 2nd ed. (1968), including bibliographies of the works of these cowinners of the Nobel Prize for Physiology or Medicine (1944). More recent works include Bertil Hille and William A. Catterall, “Electrical Excitability and Ion Channels,” in George J. Siegel (ed.), Basic Neurochemistry: Molecular, Cellular, and Medical Aspects, 6th ed. (1999), pp. 119–137, and Motoy Kuno, The Synapse: Function, Plasticity, and Neurotrophism (1995).The evolution and development of the nervous system is explored by John C. Eccles, Evolution of the Brain: Creation of the Self (1989, reissued 1995); and Richard Byrne, The Thinking Ape: Evolutionary Origins of Intelligence (1995).Thomas L. Lentz Solomon D. Erulkar

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Nervous system — Nervous Nerv ous (n[ e]rv [u^]s), a. [L. nervosus sinewy, vigorous: cf. F. nerveux. See {Nerve}.] 1. Possessing nerve; sinewy; strong; vigorous. Nervous arms. Pope. [1913 Webster] 2. Possessing or manifesting vigor of mind; characterized by… …   The Collaborative International Dictionary of English

  • nervous system — nervous systems N COUNT Your nervous system consists of all the nerves in your body together with your brain and spinal cord …   English dictionary

  • nervous system — n. all the nerve cells and nervous tissues in an organism, including, in the vertebrates, the brain, spinal cord, ganglia, nerves, and nerve centers: it coordinates and controls responses to stimuli and conditions behavior and consciousness …   English World dictionary

  • nervous system — nervous .system n your nerves, brain, and ↑spinal cord, with which your body feels pain, heat etc and your movements are controlled …   Dictionary of contemporary English

  • nervous system — ► NOUN ▪ the network of nerve cells and fibres which transmits nerve impulses between parts of the body …   English terms dictionary

  • Nervous system — Nerve redirects here. For other uses, see Nerve (disambiguation). This article is about the nervous system. For parts of the nervous system commonly called nerves, see Peripheral nerve. Nervous system The Human N …   Wikipedia

  • Nervous system — The nervous system is the body tissue that records and distributes information in the body using electrical and chemical transmission. It has two parts. The "central" nervous system is comprised of the brain and spinal cord. The… …   Medical dictionary

  • nervous system — noun the sensory and control apparatus consisting of a network of nerve cells (Freq. 2) • Syn: ↑systema nervosum • Hypernyms: ↑system • Hyponyms: ↑sympathetic nervous system, ↑parasympathetic nervous system …   Useful english dictionary

  • nervous system — /ˈnɜvəs sɪstəm/ (say nervuhs sistuhm) noun the system of nerves and nerve centres in an animal, comprising two major parts, the central nervous system and the peripheral nervous system. See central nervous system, peripheral nervous system,… …  

  • nervous system — the vast network of cells specialized to carry information (in the form of nerve impulses) to and from all parts of the body in order to bring about bodily activity. The brain and spinal cord together form the central nervous system; the… …   The new mediacal dictionary

Share the article and excerpts

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