Search This Blog

Friday, January 31, 2020

Dendritic spine

From Wikipedia, the free encyclopedia
https://en.wikipedia.org/wiki/Dendritic_spine
 
Dendritic spine
Dendritic spines.jpg
Spiny dendrite of a striatal medium spiny neuron.
Spline types 3D.png
Common types of dendritic spines.
Details
Identifiers
Latingemmula dendritica
MeSHD049229
THH2.00.06.1.00036

A dendritic spine (or spine) is a small membranous protrusion from a neuron's dendrite that typically receives input from a single axon at the synapse. Dendritic spines serve as a storage site for synaptic strength and help transmit electrical signals to the neuron's cell body. Most spines have a bulbous head (the spine head), and a thin neck that connects the head of the spine to the shaft of the dendrite. The dendrites of a single neuron can contain hundreds to thousands of spines. In addition to spines providing an anatomical substrate for memory storage and synaptic transmission, they may also serve to increase the number of possible contacts between neurons.

Structure

Dendritic spines are small with spine head volumes ranging 0.01 μm3 to 0.8 μm3. Spines with strong synaptic contacts typically have a large spine head, which connects to the dendrite via a membranous neck. The most notable classes of spine shape are "thin", "stubby", "mushroom", and "branched". Electron microscopy studies have shown that there is a continuum of shapes between these categories. The variable spine shape and volume is thought to be correlated with the strength and maturity of each spine-synapse. 

Distribution

Dendritic spines usually receive excitatory input from axons, although sometimes both inhibitory and excitatory connections are made onto the same spine head. Excitatory axon proximity to dendritic spines is not sufficient to predict the presence of a synapse, as demonstrated by the Lichtman lab in 2015.

Spines are found on the dendrites of most principal neurons in the brain, including the pyramidal neurons of the neocortex, the medium spiny neurons of the striatum, and the Purkinje cells of the cerebellum. Dendritic spines occur at a density of up to 5 spines/1 μm stretch of dendrite. Hippocampal and cortical pyramidal neurons may receive tens of thousands of mostly excitatory inputs from other neurons onto their equally numerous spines, whereas the number of spines on Purkinje neuron dendrites is an order of magnitude larger. 

Cytoskeleton and organelles

The cytoskeleton of dendritic spines is particularly important in their synaptic plasticity; without a dynamic cytoskeleton, spines would be unable to rapidly change their volumes or shapes in responses to stimuli. These changes in shape might affect the electrical properties of the spine. The cytoskeleton of dendritic spines is primarily made of filamentous actin (F-actin). tubulin Monomers and microtubule-associated proteins (MAPs) are present, and organized microtubules are present. Because spines have a cytoskeleton of primarily actin, this allows them to be highly dynamic in shape and size. The actin cytoskeleton directly determines the morphology of the spine, and actin regulators, small GTPases such as Rac, RhoA, and CDC42, rapidly modify this cytoskeleton. Overactive Rac1 results in consistently smaller dendritic spines.

In addition to their electrophysiological activity and their receptor-mediated activity, spines appear to be vesicularly active and may even translate proteins. Stacked discs of the smooth endoplasmic reticulum (SERs) have been identified in dendritic spines. Formation of this "spine apparatus" depends on the protein synaptopodin and is believed to play an important role in calcium handling. "Smooth" vesicles have also been identified in spines, supporting the vesicular activity in dendritic spines. The presence of polyribosomes in spines also suggests protein translational activity in the spine itself, not just in the dendrite. 

Physiology


Receptor activity

Dendritic spines express glutamate receptors (e.g. AMPA receptor and NMDA receptor) on their surface. The TrkB receptor for BDNF is also expressed on the spine surface, and is believed to play a role in spine survival. The tip of the spine contains an electron-dense region referred to as the "postsynaptic density" (PSD). The PSD directly apposes the active zone of its synapsing axon and comprises ~10% of the spine's membrane surface area; neurotransmitters released from the active zone bind receptors in the postsynaptic density of the spine. Half of the synapsing axons and dendritic spines are physically tethered by calcium-dependent cadherin, which forms cell-to-cell adherent junctions between two neurons. 

Glutamate receptors (GluRs) are localized to the postsynaptic density, and are anchored by cytoskeletal elements to the membrane. They are positioned directly above their signalling machinery, which is typically tethered to the underside of the plasma membrane, allowing signals transmitted by the GluRs into the cytosol to be further propagated by their nearby signalling elements to activate signal transduction cascades. The localization of signalling elements to their GluRs is particularly important in ensuring signal cascade activation, as GluRs would be unable to affect particular downstream effects without nearby signallers.

Signalling from GluRs is mediated by the presence of an abundance of proteins, especially kinases, that are localized to the postsynaptic density. These include calcium-dependent calmodulin, CaMKII (calmodulin-dependent protein kinase II), PKC (Protein Kinase C), PKA (Protein Kinase A), Protein Phosphatase-1 (PP-1), and Fyn tyrosine kinase. Certain signallers, such as CaMKII, are upregulated in response to activity.

Spines are particularly advantageous to neurons by compartmentalizing biochemical signals. This can help to encode changes in the state of an individual synapse without necessarily affecting the state of other synapses of the same neuron. The length and width of the spine neck has a large effect on the degree of compartmentalization, with thin spines being the most biochemically isolated spines. 

Plasticity

Dendritic spines are very "plastic", that is, spines change significantly in shape, volume, and number in small time courses. Because spines have a primarily actin cytoskeleton, they are dynamic, and the majority of spines change their shape within seconds to minutes because of the dynamicity of actin remodeling. Furthermore, spine number is very variable and spines come and go; in a matter of hours, 10-20% of spines can spontaneously appear or disappear on the pyramidal cells of the cerebral cortex, although the larger "mushroom"-shaped spines are the most stable.

Spine maintenance and plasticity is activity-dependent and activity-independent. BDNF partially determines spine levels, and low levels of AMPA receptor activity is necessary to maintain spine survival, and synaptic activity involving NMDA receptors encourages spine growth. Furthermore, two-photon laser scanning microscopy and confocal microscopy have shown that spine volume changes depending on the types of stimuli that are presented to a synapse. 

Importance to learning and memory


Evidence of importance

A depiction of spine formation and elimination.
Experience-dependent spine formation and elimination

Spine plasticity is implicated in motivation, learning, and memory. In particular, long-term memory is mediated in part by the growth of new dendritic spines (or the enlargement of pre-existing spines) to reinforce a particular neural pathway. Because dendritic spines are plastic structures whose lifespan is influenced by input activity, spine dynamics may play an important role in the maintenance of memory over a lifetime. 

Age-dependent changes in the rate of spine turnover suggest that spine stability impacts developmental learning. In youth, dendritic spine turnover is relatively high and produces a net loss of spines. This high rate of spine turnover may characterize critical periods of development and reflect learning capacity in adolescence—different cortical areas exhibit differing levels of synaptic turnover during development, possibly reflecting varying critical periods for specific brain regions. In adulthood, however, most spines remain persistent, and the half-life of spines increases. This stabilization occurs due to a developmentally regulated slow-down of spine elimination, a process which may underlie the stabilization of memories in maturity.

Experience-induced changes in dendritic spine stability also point to spine turnover as a mechanism involved in the maintenance of long-term memories, though it is unclear how sensory experience affects neural circuitry. Two general models might describe the impact of experience on structural plasticity. On the one hand, experience and activity may drive the discrete formation of relevant synaptic connections that store meaningful information in order to allow for learning. On the other hand, synaptic connections may be formed in excess, and experience and activity may lead to the pruning of extraneous synaptic connections.

In lab animals of all ages, environmental enrichment has been related to dendritic branching, spine density, and overall number of synapses. In addition, skill training has been shown to lead to the formation and stabilization of new spines while destabilizing old spines, suggesting that the learning of a new skill involves a rewiring process of neural circuits. Since the extent of spine remodeling correlates with success of learning, this suggests a crucial role of synaptic structural plasticity in memory formation. In addition, changes in spine stability and strengthening occur rapidly and have been observed within hours after training.

Conversely, while enrichment and training are related to increases in spine formation and stability, long-term sensory deprivation leads to an increase in the rate of spine elimination and therefore impacts long-term neural circuitry. Upon restoring sensory experience after deprivation in adolescence, spine elimination is accelerated, suggesting that experience plays an important role in the net loss of spines during development. In addition, other sensory deprivation paradigms—such as whisker trimming—have been shown to increase the stability of new spines.

Research in neurological diseases and injuries shed further light on the nature and importance of spine turnover. After stroke, a marked increase in structural plasticity occurs near the trauma site, and a five- to eightfold increase from control rates in spine turnover has been observed. Dendrites disintegrate and reassemble rapidly during ischemia—as with stroke, survivors showed an increase in dendritic spine turnover. While a net loss of spines is observed in Alzheimer's disease and cases of intellectual disability, cocaine and amphetamine use have been linked to increases in dendritic branching and spine density in the prefrontal cortex and the nucleus accumbens. Because significant changes in spine density occur in various brain diseases, this suggests a balanced state of spine dynamics in normal circumstances, which may be susceptible to disequilibrium under varying pathological conditions.

There is also some evidence for loss of dendritic spines as a consequence of aging. One study using mice has noted a correlation between age-related reductions in spine densities in the hippocampus and age-dependent declines in hippocampal learning and memory.

Importance contested

Despite experimental findings that suggest a role for dendritic spine dynamics in mediating learning and memory, the degree of structural plasticity's importance remains debatable. For instance, studies estimate that only a small portion of spines formed during training actually contribute to lifelong learning. In addition, the formation of new spines may not significantly contribute to the connectivity of the brain, and spine formation may not bear as much of an influence on memory retention as other properties of structural plasticity, such as the increase in size of spine heads.

Modeling

Theoreticians have for decades hypothesized about the potential electrical function of spines, yet our inability to examine their electrical properties has until recently stopped theoretical work from progressing too far. Recent advances in imaging techniques along with increased use of two-photon glutamate uncaging have led to a wealth of new discoveries; we now suspect that there are voltage-dependent sodium, potassium, and calcium channels in the spine heads.

Cable theory provides the theoretical framework behind the most "simple" method for modelling the flow of electrical currents along passive neural fibres. Each spine can be treated as two compartments, one representing the neck, the other representing the spine head. The compartment representing the spine head alone should carry the active properties. 

Baer and Rinzel's continuum model

To facilitate the analysis of interactions between many spines, Baer & Rinzel formulated a new cable theory for which the distribution of spines is treated as a continuum. In this representation, spine head voltage is the local spatial average of membrane potential in adjacent spines. The formulation maintains the feature that there is no direct electrical coupling between neighboring spines; voltage spread along dendrites is the only way for spines to interact.

Spike-diffuse-spike model

The SDS model was intended as a computationally simple version of the full Baer and Rinzel model. It was designed to be analytically tractable and have as few free parameters as possible while retaining those of greatest significance, such as spine neck resistance. The model drops the continuum approximation and instead uses a passive dendrite coupled to excitable spines at discrete points. Membrane dynamics in the spines are modelled using integrate and fire processes. The spike events are modelled in a discrete fashion with the wave form conventionally represented as a rectangular function. 

Modeling spine calcium transients

Calcium transients in spines are a key trigger for synaptic plasticity. NMDA receptors, which have a high permeability for calcium, only conduct ions if the membrane potential is suffiently depolarized. The amount of calcium entering a spine during synaptic activity therefore depends on the depolarization of the spine head. Evidence from calcium imaging experiments (two-photon microscopy) and from compartmental modelling indicates that spines with high resistance necks experience larger calcium transients during synaptic activity.

Development

Dendritic spines can develop directly from dendritic shafts or from dendritic filopodia. During synaptogenesis, dendrites rapidly sprout and retract filopodia, small membrane organelle-lacking membranous protrusions. Recently, I-BAR protein MIM was found to contribute to the initiation process. During the first week of birth, the brain is predominated by filopodia, which eventually develop synapses. However, after this first week, filopodia are replaced by spiny dendrites but also small, stubby spines that protrude from spiny dendrites. In the development of certain filopodia into spines, filopodia recruit presynaptic contact to the dendrite, which encourages the production of spines to handle specialized postsynaptic contact with the presynaptic protrusions. 

Spines, however, require maturation after formation. Immature spines have impaired signaling capabilities, and typically lack "heads" (or have very small heads), only necks, while matured spines maintain both heads and necks.

Clinical significance

Cognitive disorders such as ADHD, autism, intellectual disability, and fragile X syndrome, may be resultant from abnormalities in dendritic spines, especially the number of spines and their maturity. The ratio of matured to immature spines is important in their signaling, as immature spines have impaired synaptic signaling. Fragile X syndrome is characterized by an overabundance of immature spines that have multiple filopodia in cortical dendrites. 

History

Dendritic spines were first described at the end of the 19th century by Santiago Ramón y Cajal on cerebellar neurons. Ramón y Cajal then proposed that dendritic spines could serve as contacting sites between neurons. This was demonstrated more than 50 years later thanks to the emergence of electron microscopy. Until the development of confocal microscopy on living tissues, it was commonly admitted that spines were formed during embryonic development and then would remain stable after birth. In this paradigm, variations of synaptic weight were considered as sufficient to explain memory processes at the cellular level. But since about a decade ago, new techniques of confocal microscopy demonstrated that dendritic spines are indeed motile and dynamic structures that undergo a constant turnover, even after birth.

Chemical synapse

From Wikipedia, the free encyclopedia
https://en.wikipedia.org/wiki/Chemical_synapse
 
Artistic interpretation of the major elements in chemical synaptic transmission. An electrochemical wave called an action potential travels along the axon of a neuron. When the action potential reaches the presynaptic terminal, it provokes the release of a synaptic vesicle, secreting its quanta of neurotransmitter molecules. The neurotransmitter binds to chemical receptor molecules located in the membrane of another neuron, the postsynaptic neuron, on the opposite side of the synaptic cleft.

Chemical synapses are biological junctions through which neurons' signals can be sent to each other and to non-neuronal cells such as those in muscles or glands. Chemical synapses allow neurons to form circuits within the central nervous system. They are crucial to the biological computations that underlie perception and thought. They allow the nervous system to connect to and control other systems of the body. 

At a chemical synapse, one neuron releases neurotransmitter molecules into a small space (the synaptic cleft) that is adjacent to another neuron. The neurotransmitters are contained within small sacs called synaptic vesicles, and are released into the synaptic cleft by exocytosis. These molecules then bind to neurotransmitter receptors on the postsynaptic cell. Finally, the neurotransmitters are cleared from the synapse through one of several potential mechanisms including enzymatic degradation or re-uptake by specific transporters either on the presynaptic cell or on some other neuroglia to terminate the action of the neurotransmitter. 

The adult human brain is estimated to contain from 1014 to 5 × 1014 (100–500 trillion) synapses. Every cubic millimeter of cerebral cortex contains roughly a billion (short scale, i.e. 109) of them. The number of synapses in the human cerebral cortex has separately been estimated at 0.15 quadrillion (150 trillion).
 
The word "synapse" comes from "synaptein", which Sir Charles Scott Sherrington and colleagues coined from the Greek "syn-" ("together") and "haptein" ("to clasp"). Chemical synapses are not the only type of biological synapse: electrical and immunological synapses also exist. Without a qualifier, however, "synapse" commonly refers to chemical synapse.

Structure

Structure of a typical chemical synapse
Distinguish between pre- and post- synapse
"The connection linking neuron to neuron is the synapse. Signal flows
in one direction, from the presynaptic neuron to the postsynaptic neuron
via the synapse which acts as a variable attenuator."  In brief,
the direction of the signal flow determines the prefix for the involved
synapses.

Synapses are functional connections between neurons, or between neurons and other types of cells. A typical neuron gives rise to several thousand synapses, although there are some types that make far fewer. Most synapses connect axons to dendrites, but there are also other types of connections, including axon-to-cell-body, axon-to-axon, and dendrite-to-dendrite. Synapses are generally too small to be recognizable using a light microscope except as points where the membranes of two cells appear to touch, but their cellular elements can be visualized clearly using an electron microscope.

Chemical synapses pass information directionally from a presynaptic cell to a postsynaptic cell and are therefore asymmetric in structure and function. The presynaptic axon terminal, or synaptic bouton, is a specialized area within the axon of the presynaptic cell that contains neurotransmitters enclosed in small membrane-bound spheres called synaptic vesicles (as well as a number of other supporting structures and organelles, such as mitochondria and endoplasmic reticulum). Synaptic vesicles are docked at the presynaptic plasma membrane at regions called active zones

Immediately opposite is a region of the postsynaptic cell containing neurotransmitter receptors; for synapses between two neurons the postsynaptic region may be found on the dendrites or cell body. Immediately behind the postsynaptic membrane is an elaborate complex of interlinked proteins called the postsynaptic density (PSD). 

Proteins in the PSD are involved in anchoring and trafficking neurotransmitter receptors and modulating the activity of these receptors. The receptors and PSDs are often found in specialized protrusions from the main dendritic shaft called dendritic spines.

Synapses may be described as symmetric or asymmetric. When examined under an electron microscope, asymmetric synapses are characterized by rounded vesicles in the presynaptic cell, and a prominent postsynaptic density. Asymmetric synapses are typically excitatory. Symmetric synapses in contrast have flattened or elongated vesicles, and do not contain a prominent postsynaptic density. Symmetric synapses are typically inhibitory. 

The synaptic cleft —also called synaptic gap— is a gap between the pre- and postsynaptic cells that is about 20 nm (0.02 μ) wide. The small volume of the cleft allows neurotransmitter concentration to be raised and lowered rapidly.

An autapse is a chemical (or electrical) synapse formed when the axon of one neuron synapses with its own dendrites. 

Signaling in chemical synapses


Overview

Here is a summary of the sequence of events that take place in synaptic transmission from a presynaptic neuron to a postsynaptic cell. Each step is explained in more detail below. Note that with the exception of the final step, the entire process may run only a few hundred microseconds, in the fastest synapses.

  1. The process begins with a wave of electrochemical excitation called an action potential traveling along the membrane of the presynaptic cell, until it reaches the synapse.
  2. The electrical depolarization of the membrane at the synapse causes channels to open that are permeable to calcium ions.
  3. Calcium ions flow through the presynaptic membrane, rapidly increasing the calcium concentration in the interior.
  4. The high calcium concentration activates a set of calcium-sensitive proteins attached to vesicles that contain a neurotransmitter chemical.
  5. These proteins change shape, causing the membranes of some "docked" vesicles to fuse with the membrane of the presynaptic cell, thereby opening the vesicles and dumping their neurotransmitter contents into the synaptic cleft, the narrow space between the membranes of the pre- and postsynaptic cells.
  6. The neurotransmitter diffuses within the cleft. Some of it escapes, but some of it binds to chemical receptor molecules located on the membrane of the postsynaptic cell.
  7. The binding of neurotransmitter causes the receptor molecule to be activated in some way. Several types of activation are possible, as described in more detail below. In any case, this is the key step by which the synaptic process affects the behavior of the postsynaptic cell.
  8. Due to thermal vibration, the motion of atoms, vibrating about their equilibrium positions in a crystalline solid, neurotransmitter molecules eventually break loose from the receptors and drift away.
  9. The neurotransmitter is either reabsorbed by the presynaptic cell, and then repackaged for future release, or else it is broken down metabolically.

Neurotransmitter release

Release of neurotransmitter occurs at the end of axonal branches.
 
The release of a neurotransmitter is triggered by the arrival of a nerve impulse (or action potential) and occurs through an unusually rapid process of cellular secretion (exocytosis). Within the presynaptic nerve terminal, vesicles containing neurotransmitter are localized near the synaptic membrane. The arriving action potential produces an influx of calcium ions through voltage-dependent, calcium-selective ion channels at the down stroke of the action potential (tail current). Calcium ions then bind to synaptotagmin proteins found within the membranes of the synaptic vesicles, allowing the vesicles to fuse with the presynaptic membrane. The fusion of a vesicle is a stochastic process, leading to frequent failure of synaptic transmission at the very small synapses that are typical for the central nervous system. Large chemical synapses (e.g. the neuromuscular junction), on the other hand, have a synaptic release probability of 1. Vesicle fusion is driven by the action of a set of proteins in the presynaptic terminal known as SNAREs. As a whole, the protein complex or structure that mediates the docking and fusion of presynaptic vesicles is called the active zone. The membrane added by the fusion process is later retrieved by endocytosis and recycled for the formation of fresh neurotransmitter-filled vesicles. 

An exception to the general trend of neurotransmitter release by vesicular fusion is found in the type II receptor cells of mammalian taste buds. Here the neurotransmitter ATP is released directly from the cytoplasm into the synaptic cleft via voltage gated channels.

Receptor binding

Receptors on the opposite side of the synaptic gap bind neurotransmitter molecules. Receptors can respond in either of two general ways. First, the receptors may directly open ligand-gated ion channels in the postsynaptic cell membrane, causing ions to enter or exit the cell and changing the local transmembrane potential. The resulting change in voltage is called a postsynaptic potential. In general, the result is excitatory in the case of depolarizing currents, and inhibitory in the case of hyperpolarizing currents. Whether a synapse is excitatory or inhibitory depends on what type(s) of ion channel conduct the postsynaptic current(s), which in turn is a function of the type of receptors and neurotransmitter employed at the synapse. The second way a receptor can affect membrane potential is by modulating the production of chemical messengers inside the postsynaptic neuron. These second messengers can then amplify the inhibitory or excitatory response to neurotransmitters.

Termination

After a neurotransmitter molecule binds to a receptor molecule, it must be removed to allow for the postsynaptic membrane to continue to relay subsequent EPSPs and/or IPSPs. This removal can happen through one or more processes:
  • The neurotransmitter may diffuse away due to thermally-induced oscillations of both it and the receptor, making it available to be broken down metabolically outside the neuron or to be reabsorbed.
  • Enzymes within the subsynaptic membrane may inactivate/metabolize the neurotransmitter.
  • Reuptake pumps may actively pump the neurotransmitter back into the presynaptic axon terminal for reprocessing and re-release following a later action potential.

Synaptic strength

The strength of a synapse has been defined by Sir Bernard Katz as the product of (presynaptic) release probability pr, quantal size q (the postsynaptic response to the release of a single neurotransmitter vesicle, a 'quantum'), and n, the number of release sites. "Unitary connection" usually refers to an unknown number of individual synapses connecting a presynaptic neuron to a postsynaptic neuron. The amplitude of postsynaptic potentials (PSPs) can be as low as 0.4 mV to as high as 20 mV. The amplitude of a PSP can be modulated by neuromodulators or can change as a result of previous activity. Changes in the synaptic strength can be short-term, lasting seconds to minutes, or long-term (long-term potentiation, or LTP), lasting hours. Learning and memory are believed to result from long-term changes in synaptic strength, via a mechanism known as synaptic plasticity

Receptor desensitization

Desensitization of the postsynaptic receptors is a decrease in response to the same neurotransmitter stimulus. It means that the strength of a synapse may in effect diminish as a train of action potentials arrive in rapid succession – a phenomenon that gives rise to the so-called frequency dependence of synapses. The nervous system exploits this property for computational purposes, and can tune its synapses through such means as phosphorylation of the proteins involved.

Synaptic plasticity

Synaptic transmission can be changed by previous activity. These changes are called synaptic plasticity and may result in either a decrease in the efficacy of the synapse, called depression, or an increase in efficacy, called potentiation. These changes can either be long-term or short-term. Forms of short-term plasticity include synaptic fatigue or depression and synaptic augmentation. Forms of long-term plasticity include long-term depression and long-term potentiation. Synaptic plasticity can be either homosynaptic (occurring at a single synapse) or heterosynaptic (occurring at multiple synapses).

Homosynaptic plasticity

Homosynaptic Plasticity (or also homotropic modulation) is a change in the synaptic strength that results from the history of activity at a particular synapse. This can result from changes in presynaptic calcium as well as feedback onto presynaptic receptors, i.e. a form of autocrine signaling. Homosynaptic plasticity can affect the number and replenishment rate of vesicles or it can affect the relationship between calcium and vesicle release. Homosynaptic plasticity can also be postsynaptic in nature. It can result in either an increase or decrease in synaptic strength.

One example is neurons of the sympathetic nervous system (SNS), which release noradrenaline, which, besides affecting postsynaptic receptors, also affects presynaptic α2-adrenergic receptors, inhibiting further release of noradrenaline. This effect is utilized with clonidine to perform inhibitory effects on the SNS.

Heterosynaptic plasticity

Heterosynaptic Plasticity (or also heterotropic modulation) is a change in synaptic strength that results from the activity of other neurons. Again, the plasticity can alter the number of vesicles or their replenishment rate or the relationship between calcium and vesicle release. Additionally, it could directly affect calcium influx. Heterosynaptic plasticity can also be postsynaptic in nature, affecting receptor sensitivity. 

One example is again neurons of the sympathetic nervous system, which release noradrenaline, which, in addition, generates an inhibitory effect on presynaptic terminals of neurons of the parasympathetic nervous system.

Integration of synaptic inputs

In general, if an excitatory synapse is strong enough, an action potential in the presynaptic neuron will trigger an action potential in the postsynaptic cell. In many cases the excitatory postsynaptic potential (EPSP) will not reach the threshold for eliciting an action potential. When action potentials from multiple presynaptic neurons fire simultaneously, or if a single presynaptic neuron fires at a high enough frequency, the EPSPs can overlap and summate. If enough EPSPs overlap, the summated EPSP can reach the threshold for initiating an action potential. This process is known as summation, and can serve as a high pass filter for neurons.

On the other hand, a presynaptic neuron releasing an inhibitory neurotransmitter, such as GABA, can cause an inhibitory postsynaptic potential (IPSP) in the postsynaptic neuron, bringing the membrane potential farther away from the threshold, decreasing its excitability and making it more difficult for the neuron to initiate an action potential. If an IPSP overlaps with an EPSP, the IPSP can in many cases prevent the neuron from firing an action potential. In this way, the output of a neuron may depend on the input of many different neurons, each of which may have a different degree of influence, depending on the strength and type of synapse with that neuron. John Carew Eccles performed some of the important early experiments on synaptic integration, for which he received the Nobel Prize for Physiology or Medicine in 1963. 

Volume transmission

When a neurotransmitter is released at a synapse, it reaches its highest concentration inside the narrow space of the synaptic cleft, but some of it is certain to diffuse away before being reabsorbed or broken down. If it diffuses away, it has the potential to activate receptors that are located either at other synapses or on the membrane away from any synapse. The extrasynaptic activity of a neurotransmitter is known as volume transmission. It is well established that such effects occur to some degree, but their functional importance has long been a matter of controversy.

Recent work indicates that volume transmission may be the predominant mode of interaction for some special types of neurons. In the mammalian cerebral cortex, a class of neurons called neurogliaform cells can inhibit other nearby cortical neurons by releasing the neurotransmitter GABA into the extracellular space. Along the same vein, GABA released from neurogliaform cells into the extracellular space also acts on surrounding astrocytes, assigning a role for volume transmission in the control of ionic and neurotransmitter homeostasis. Approximately 78% of neurogliaform cell boutons do not form classical synapses. This may be the first definitive example of neurons communicating chemically where classical synapses are not present.

Relationship to electrical synapses

An electrical synapse is an electrically conductive link between two abutting neurons that is formed at a narrow gap between the pre- and postsynaptic cells, known as a gap junction. At gap junctions, cells approach within about 3.5 nm of each other, rather than the 20 to 40 nm distance that separates cells at chemical synapses. As opposed to chemical synapses, the postsynaptic potential in electrical synapses is not caused by the opening of ion channels by chemical transmitters, but rather by direct electrical coupling between both neurons. Electrical synapses are faster than chemical synapses. Electrical synapses are found throughout the nervous system, including in the retina, the reticular nucleus of the thalamus, the neocortex, and in the hippocampus. While chemical synapses are found between both excitatory and inhibitory neurons, electrical synapses are most commonly found between smaller local inhibitory neurons. Electrical synapses can exist between two axons, two dendrites, or between an axon and a dendrite. In some fish and amphibians, electrical synapses can be found within the same terminal of a chemical synapse, as in Mauthner cells.

Effects of drugs

One of the most important features of chemical synapses is that they are the site of action for the majority of psychoactive drugs. Synapses are affected by drugs such as curare, strychnine, cocaine, morphine, alcohol, LSD, and countless others. These drugs have different effects on synaptic function, and often are restricted to synapses that use a specific neurotransmitter. For example, curare is a poison that stops acetylcholine from depolarizing the postsynaptic membrane, causing paralysis. Strychnine blocks the inhibitory effects of the neurotransmitter glycine, which causes the body to pick up and react to weaker and previously ignored stimuli, resulting in uncontrollable muscle spasms. Morphine acts on synapses that use endorphin neurotransmitters, and alcohol increases the inhibitory effects of the neurotransmitter GABA. LSD interferes with synapses that use the neurotransmitter serotonin. Cocaine blocks reuptake of dopamine and therefore increases its effects.

History

During the 1950s, Bernard Katz and Paul Fatt observed spontaneous miniature synaptic currents at the frog neuromuscular junction. Based on these observations, they developed the 'quantal hypothesis' that is the basis for our current understanding of neurotransmitter release as exocytosis and for which Katz received the Nobel Prize in Physiology or Medicine in 1970. In the late 1960s, Ricardo Miledi and Katz advanced the hypothesis that depolarization-induced influx of calcium ions triggers exocytosis.

Activity-dependent plasticity

From Wikipedia, the free encyclopedia
https://en.wikipedia.org/wiki/Activity-dependent_plasticity

Activity-dependent plasticity is a form of functional and structural neuroplasticity that arises from the use of cognitive functions and personal experience; hence, it is the biological basis for learning and the formation of new memories. Activity-dependent plasticity is a form of neuroplasticity that arises from intrinsic or endogenous activity, as opposed to forms of neuroplasticity that arise from extrinsic or exogenous factors, such as electrical brain stimulation- or drug-induced neuroplasticity. The brain's ability to remodel itself forms the basis of the brain's capacity to retain memories, improve motor function, and enhance comprehension and speech amongst other things. It is this trait to retain and form memories that is associated with neural plasticity and therefore many of the functions individuals perform on a daily basis. This plasticity occurs as a result of changes in gene expression which are triggered by signaling cascades that are activated by various signaling molecules (e.g., calcium, dopamine, and glutamate, among many others) during increased neuronal activity.

The brain's ability to adapt toward active functions allows humans to specialize in specific processes based on relative use and activity. For example, a right-handed person may perform any movement poorly with his/her left hand but continuous practice with the less dominant hand can cause one to become ambidextrous. Another example is if someone was born with a neurological disorder such as major depressive disorder or had a stroke that resulted in a disorder, then they are capable of retrieving much of their lost function through practice, which in turn "rewires" the brain to mitigate neurological dysfunction.

History

The idea of neural plasticity was first proposed during 1890 by William James in Principles of Psychology. During the first half of the 1900s, the word 'plasticity' was directly and indirectly rejected throughout science. Many scientists found it hard to receive funding because nearly everyone unanimously supported the fact that the brain was fully developed at adulthood and specific regions were unable to change functions after the critical period. It was believed that each region of the brain had a set and specific function. Despite this, several pioneers pushed the idea of plasticity through means of various experiments and research. There are others that helped to the current progress of activity-dependent plasticity but the following contributed very effective results and ideas early on. 

Pioneers of activity-dependent plasticity

The history of activity-dependent plasticity begins with Paul Bach y Rita. With conventional ideology being that the brain development is finalized upon adulthood, Bach y Rita designed several experiments in the late 1960s and 1970s that proved that the brain is capable of changing. These included a pivotal visual substitution method for blind people provided by tactile image projection in 1969. The basis behind this experiment was to take one sense and use it to detect another: in this case use the sense of touch on the tongue to visualize the surrounding. This experiment was years ahead of its time and lead to many questions and applications. A similar experiment was reported again by Bach y Rita in 1986 where vibrotactile stimulation was delivered to the index fingertips of naive blindfolded subjects. Even though the experiment did not yield great results, it supported the study and proposed further investigations. In 1998, his design was even further developed and tested again with a 49-point electrotactile stimulus array on the tongue. He found that five sighted adult subjects recognized shapes across all sizes 79.8% of the time, a remarkable finding that has led to the incorporation of the tongue electrotactile stimulus into cosmetically acceptable and practical designs for blind people. In later years, he has published a number of other articles including "Seeing with the brain" in 2003 where Bach y Rita addresses the plasticity of the brain relative to visual learning. Here, images are enhanced and perceived by other plastic mechanisms within the realm of information passing to the brain. 

Another pioneer within the field of activity-dependent plasticity is Michael Merzenich, currently a professor in neuroscience at the University of California, San Francisco. One of his contributions includes mapping out and documenting the reorganization of cortical regions after alterations due to plasticity. While assessing the recorded changes in the primary somatosensory cortex of adult monkeys, he looked at several features of the data including how altered schedules of activity from the skin remap to cortical modeling and other factors that affect the representational remodeling of the brain. His findings within these studies have since been applied to youth development and children with language-based learning impairments. Through many studies involving adaptive training exercises on computer, he has successfully designed methods to improve their temporal processing skills. These adaptive measures include word-processing games and comprehension tests that involve multiple regions of the brain in order to answer. The results later translated into his development of the Fast ForWord program in 1996, which aims to enhance cognitive skills of children between kindergarten and twelfth grade by focusing on developing "phonological awareness". It has proven very successful at helping children with a variety of cognitive complications. In addition, it has led to in depth studies of specific complications such as autism and intellectual disability and the causes of them. Alongside a team of scientists, Merzenich helped to provide evidence that autism probes monochannel perception where a stronger stimulus-driven representation dominates behavior and weaker stimuli are practically ignored in comparison. 

Structure of neurons

Diagram displaying components of a myelinated vertebrate motorneuron.

Neurons are the basic functional unit of the brain and process and transmit information through signals. Many different types of neurons can be identified based on their function, such as sensory neurons or motor neurons. Each responds to specific stimuli and sends respective and appropriate chemical signals to other neurons. The basic structure of a neuron is shown here on the right and consists of a nucleus that contains genetic information; the cell body, or the soma, which is equipped with dendritic branches that mostly receive the incoming inputs from other neurons; a long, thin axon that bears axon terminals which carry the output information to other neurons. The dendrites and axons are interfaced through a small connection called a synapse. This component of the neuron contains a variety of chemical messengers and proteins that allow for the transmission of information. It is the variety of proteins and effect of the signal that fundamentally lead to the plasticity feature. 

Structures and molecular pathways involved

Activity-dependent plasticity of one form or another has been observed in most areas of the brain. In particular, it is thought that the reorganization of sensory and motor maps involves a variety of pathways and cellular structures related to relative activity.

Many molecules have been implicated in synaptic plasticity. Notably, AMPA and NMDA receptors are key molecules in mechanisms of long and short-term potentiation between neurons. NMDA receptors can detect local activity due to activation and therefore modify signaling in the post-synaptic cell. The increased activity and coordination between pre- and post-synaptic receptors leads to more permanent changes and therefore result in plasticity. Hebb's postulate addresses this fact by stating that synaptic terminals are strengthened by correlated activity and will therefore sprout new branches. However, terminals that experience weakened and minimal activity will eventually lose their synaptic connection and deteriorate.

A major target of all molecular signaling is the inhibitory connections made by GABAergic neurons. These receptors exist at postsynaptic sites and along with the regulation of local inhibitory synapses have been found to be very sensitive to critical period alterations. Any alteration to the receptors leads to changed concentrations of calcium in the affected cells and can ultimately influence dendritic and axonal branching. This concentration change is the result of many kinases being activated, the byproduct of which may enhance specific gene expression.

Illustration of the elements incorporated in synaptic transmission. An action potential is generated and travels down the axon to the axon terminal, where it is released and provokes a neurotransmitter release that acts on the post-synaptic end.
 
In addition, it has been identified that the wg postsynaptic pathway, which is responsible for the coding and production of many molecules for development events, can be bidirectionally stimulated and is responsible for the downstream alteration of the postsynaptic neuron. When the wg presynaptic pathway is activated, however, it alters cytoskeletal structure through transcription and translation.

Cell adhesion molecules (CAMs) are also important in plasticity as they help coordinate the signaling across the synapse. More specifically, integrins, which are receptors for extracellular matrix proteins and involved with CAMs, are explicitly incorporated in synapse maturation and memory formation. They play a crucial role in the feedback regulation of excitatory synaptic strength, or long-term potentiation (LTP), and help to control synaptic strength by regulating AMPA receptors, which result in quick, short synaptic currents. But, it is the metabotropic glutamate receptor 1 (mGlu1) that has been discovered to be required for activity-dependent synaptic plasticity in associative learning.

Activity-dependent plasticity is seen in the primary visual cortex, a region of the brain that processes visual stimuli and is capable of modifying the experienced stimuli based on active sensing and arousal states. It is known that synaptic communication trends between excited and depressed states relative to the light/dark cycle. By experimentation on rats, it was found that visual experience during vigilant states leads to increased responsiveness and plastic changes in the visual cortex. More so, depressed states were found to negatively alter the stimulus so the reaction was not as energetic. This experiment proves that even the visual cortex is capable of achieving activity-dependent plasticity as it is reliant on both visual exploration and the arousal state of the animal.

Role in learning

Activity-dependent plasticity plays a very important role in learning and in the ability of understanding new things. It is responsible for helping to adapt an individual's brain according to the relative amount of usage and functioning. In essence, it is the brain's ability to retain and develop memories based on activity-driven changes of synaptic strength that allow stronger learning of information. It is thought to be the growing and adapting quality of dendritic spines that provide the basis for synaptic plasticity connected to learning and memory. Dendritic spines accomplish this by transforming synaptic input into neuronal output and also by helping to define the relationship between synapses. 

In recent studies, a specific gene has also been identified as having a strong role in synapse growth and activity-dependent plasticity: the microRNA 132 gene (miR132). This gene is regulated by the cAMP response element-binding (CREB) protein pathway and is capable of enhancing dendritic growth when activated. The miR132 gene is another component that is responsible for the brain's plasticity and helps to establish stronger connections between neurons. 

Another plasticity-related gene involved in learning and memory is Arc/Arg3.1. The Arc gene is activity-regulated and the transcribed mRNA is localized to activated synaptic sites where the translated protein plays a role in AMPA receptor trafficking. Arc is a member of a class of proteins called immediate early genes (IEG) that are rapidly transcribed in response to synaptic input. Of the estimated 30-40 genes that comprise the total neuronal IEG response, all are prototypical activity-dependent genes and a number have been implicated in learning and memory. For example, zif268, Arc, beta-activin, tPA, Homer, and COX-2 have all been implicated in long-term potentiation (LTP), a cellular correlate of learning and memory. 

Mechanisms involved

There are a variety of mechanisms involved in activity-dependent plasticity. These include LTP, long-term depression (LTD), synaptic elimination, neurogenesis, and synaptogenesis. The mechanisms of activity-dependent plasticity result in membrane depolarization and calcium influx, which in turn trigger cellular changes that affect synaptic connections and gene transcription. In essence, neuronal activity regulates gene expression related to dendritic branching and synapse development. Mutations in activity-dependent transcription-related genes can lead to neurological disorders. Each of the studies' findings aims to help proper development of the brain while improving a wide variety of tasks such as speech, movement, comprehension, and memory. More so, the findings better explain the development induced by plasticity. 

It is known that during postnatal life a critical step to nervous system development is synapse elimination. The changes in synaptic connections and strength are results from LTP and LTD and are strongly regulated by the release of brain-derived neurotrophic factor (BDNF), an activity-dependent synapse-development protein. In addition to BDNF, Nogo-66 receptors, and more specifically NgR1, are also involved in the development and regulation of neuronal structure. Damage to this receptor leads to pointless LTP and attenuation of LTD. Both situations imply that NgR1 is a regulator of synaptic plasticity. From experiments, it has been found that stimulation inducing LTD leads to a reduction in synaptic strength and loss of connections but, when coupled simultaneously with low-frequency stimulation, helps the restructuring of synaptic contacts. The implications of this finding include helping people with receptor damage and providing insight into the mechanism behind LTP.

Another research model of activity-dependent plasticity includes the excitatory corticostriatal pathway that is involved in information processing related to adaptive motor behaviors and displays long-lasting synaptic changes. The change in synaptic strength is responsible for motor learning and is dependent on the simultaneous activation of glutamatergic corticostriatal and dopaminergic nigrostriatal pathways. These are the same pathways affected in Parkinson's disease, and the degeneration of synapses within this disorder may be responsible for the loss of some cognitive abilities.

Relationship to behavior


Intellectual disability

Since plasticity is such a fundamental property of brain function due to its involvement in brain development, brain repair, and cognitive processes, its proper regulation is necessary for normal physiology. Mutations within any of the genes associated with activity-dependent plasticity have been found to positively correlate with various degrees of intellectual disability. The two types of intellectual disability related to plasticity depend on dysfunctional neuronal development or alterations in molecular mechanisms involved in synaptic organization. Complications within either of these types can greatly reduce brain capability and comprehension.

Stroke rehabilitation

On the other hand, people with such conditions have the capacity to recover some degree of their lost abilities through continued challenges and use. An example of this can be seen in Norman Doidge's The Brain That Changes Itself. Bach y Rita's father suffered from a disabling stroke that left the 65-year-old man half-paralyzed and unable to speak. After one year of crawling and unusual therapy tactics including playing basic children's games and washing pots, his father's rehabilitation was nearly complete and he went back to his role as a professor at City College in New York. This remarkable recovery from a stroke proves that even someone with abnormal behavior and severe medical complications can recover nearly all of the normal functions by much practice and perseverance. 

Recent studies have reported that a specific gene, FMR1, is highly involved in activity-dependent plasticity and fragile X syndrome (FraX) is the result of this gene's loss of function. FMR1 produces FMRP, which mediates activity-dependent control of synaptic structure. The loss or absence of this gene almost certainly leads to both autism and intellectual disability. Dr. Gatto has found that early introduction of the product FMRP results in nearly complete restructuring of the synapses. This method is not as effective, though, when introduced into a mature subject and only partially accommodates for the losses of FMR1. The discovery of this gene provides a possible location for intervention for young children with these abnormalities as this gene and its product act early to construct synaptic architecture.

Stress

A common issue amongst most people in the United States is high levels of stress and also disorders associated with continuous stress. Many regions of the brain are very sensitive to stress and can be damaged with extended exposure. More importantly, many of the mechanisms involved with increased memory retention, comprehension, and adaptation are thought to involve LTP and LTD, two activity-dependent plasticity mechanisms that stress can directly suppress. Several experiments have been conducted in order to discover the specific mechanisms for this suppression and also possible intervention methods. Dr. Li and several others have actually identified the TRPV1 channel as a target to facilitate LTP and suppress LTD, therefore helping to protect the feature of synaptic plasticity and retention of memory from the effects of stress.

Future studies

The future studies and questions for activity-dependent plasticity are nearly endless because the implications of the findings will enable many treatments. Despite many gains within the field, there are a wide variety of disorders that further understanding of activity-dependent mechanisms of plasticity would help treat and perhaps cure. These include autism, different severities of intellectual disability, schizophrenia, Parkinson's disease, stress, and stroke. In addition to a better understanding of the various disorders, neurologists should and will look at the plasticity incurred by the immune system, as it will provide great insight into diseases and also give the basis of new immune-centered therapeutics. A better perspective of the cellular mechanisms that regulate neuronal morphology is the next step to discovering new treatments for learning and memory pathological conditions.

Neuroplastic effects of pollution

From Wikipedia, the free encyclopedia
https://en.wikipedia.org/wiki/Neuroplastic_effects_of_pollution

Research indicates that living in areas of high pollution has serious long term health effects. Living in these areas during childhood and adolescence can lead to diminished mental capacity and an increased risk of brain damage. People of all ages who live in high pollution areas for extended periods place themselves at increased risk of various neurological disorders. Both air pollution and heavy metal pollution have been implicated as having negative effects on central nervous system (CNS) functionality. The ability of pollutants to affect the neurophysiology of individuals after the structure of the CNS has become mostly stabilized is an example of negative neuroplasticity.

Air pollution

Air pollution is known to affect small and large blood vessels throughout the body. High levels of air pollution are associated with increased risk of strokes and heart attacks. By permanently affecting vascular structures in the brain, air pollution can have serious effects on neural functioning and neural matter. In dogs air pollution shows to cause damage to the CNS by altering the blood–brain barrier, causing neurons in the cerebral cortex to degenerate, destroying glial cells found in white matter, and by causing neurofibrillary tangles. These changes can permanently alter brain structure and chemistry, resulting in various impairments and disorders. Sometimes, the effects of neural remodeling do not manifest themselves for a prolonged period of time.

Effects in adolescents and canines

A study from 2008 compared children and dogs raised in Mexico City (a location known for high pollution levels) with children and dogs raised in Polotitlán, Mexico (a city whose pollution levels meet the current US National Ambient Air Quality Standards). According to this study, children raised in areas of higher pollution scored lower in intelligence (i.e. on IQ tests), and showed signs of lesions in MRI scanning of the brain. In contrast, children from the low pollution area scored as expected on IQ tests, and did not show any significant sign of the risk of brain lesions. This correlation was found to be statistically significant, and shows that pollution levels may be related to, and contribute to, brain lesion formation and IQ scores, which, in turn, manifests as impaired intellectual capacity and/or performance. Living in high pollution areas thus places adolescents at risk of premature brain degeneration and improper neural development—these findings could have significant implications for future generations.

Effects in adults

There are indications that the effects of physical activity and air pollution on neuroplasticity counteract. Physical activity is known for its health-enhancing benefits, particularly on the cardiovascular system, and has also demonstrated benefits for brain plasticity processes, cognition and mental health. The neurotrophine, brain-derived neurotrophic factor (BDNF) is thought to play a key role in exercise-induced cognitive improvements. Brief bouts of physical activity have been shown to increase serum levels of BDNF, but this increase may be offset by increased exposure to traffic-related air pollution. Over longer periods of physical exercise, cognitive improvements that were demonstrated in rural joggers were found to be absent in urban joggers taking the same 12-week start-2-run training programme.

Epilepsy

Researchers in Chile found statistically-significant correlations between multiple air pollutants and the risk of epilepsy using a 95% confidence interval. The air pollutants that the researchers attempted to correlate with increased incidence of epilepsy included carbon monoxide, ozone, sulfur dioxide, nitrogen dioxide, large particulate matter, and fine particulate matter. The researchers tested these pollutants across seven cities and, in all but one case, a correlation was found between pollutant levels and the occurrence of epilepsy. All of the correlations found were shown to be statistically significant. The researchers hypothesized that air pollutants increase epilepsy risk by increasing inflammatory mediators, and by providing a source of oxidative stress. They believe that these changes eventually alter the functioning of the blood–brain barrier, causing brain inflammation. Brain inflammation is known to be a risk factor for epilepsy; thus, the sequence of events provides a plausible mechanism by which pollution may increase epilepsy risk in individuals who are genetically vulnerable to the disease. 

Dioxin poisoning

Organohalogen compounds, such as dioxins, are commonly found in pesticides or created as by-products of pesticide manufacture or degradation. These compounds can have a significant impact on the neurobiology of exposed organisms. Some observed effects of exposure to dioxins are altered astroglial intracellular calcium ion (Ca2+), decreased glutathione levels, modified neurotransmitter function in the CNS, and loss of pH maintenance. A study of 350 chemical plant employees exposed to a dioxin precursor for herbicide synthesis between 1965 and 1968 showed that 80 of the employees displayed signs of dioxin poisoning. Of these 350 employees, 15 were contacted again in 2004 to submit to neurological tests to assess whether the dioxin poisoning had any long-term effects on neurological capabilities. The amount of time that had passed made it difficult to assemble a larger cohort, but the results of the tests indicated that eight of the 15 subjects exhibited some central nervous system impairment, nine showed signs of polyneuropathy, and electroencephalography (EEG) showed various degrees of structural abnormalities. This study suggested that the effects of dioxins were not limited to initial toxicity. Dioxins, through neuroplastic effects, can cause long-term damage that may not manifest itself for years or even decades. 

Metal exposure

Heavy metal exposure can result in an increased risk of various neurological diseases. Research indicates that the two most neurotoxic heavy metals are mercury and lead. The impact that these two metals will have is highly dependent upon the individual due to genetic variations. Mercury and lead are particularly neurotoxic for many reasons: they easily cross cell membranes, have oxidative effects on cells, react with sulfur in the body (leading to disturbances in the many functions that rely upon sulfhydryl groups), and reduce glutathione levels inside cells. Methylmercury, in particular, has an extremely high affinity for sulfhydryl groups. Organomercury is a particularly damaging form of mercury because of its high absorbability Lead also mimics calcium, a very important mineral in the CNS, and this mimicry leads to many adverse effects. Mercury's neuroplastic mechanisms work by affecting protein production. Elevated mercury levels increase glutathione levels by affecting gene expression, and this in turn affects two proteins (MT1 and MT2) that are contained in astrocytes and neurons. Lead's ability to imitate calcium allows it to cross the blood–brain barrier. Lead also upregulates glutathione.

Autism

Heavy metal exposure, when combined with certain genetic predispositions, can place individuals at increased risk for developing autism. Many examples of CNS pathophysiology, such as oxidative stress, neuroinflammation, and mitochondrial dysfunction, could be by-products of environmental stressors such as pollution. There have been reports of autism outbreaks occurring in specific locations. Since these cases of autism are related to geographic location, the implication is that something in the environment is complementing an at-risk genotype to cause autism in these vulnerable individuals. Mercury and lead both contribute to inflammation, leading scientists to speculate that these heavy metals could play a role in autism. These findings are controversial, however, with many researchers believing that increasing rates of autism are a consequence of more accurate screening and diagnostic methods, and are not due to any sort of environmental factor.

Accelerated neural aging

Neuroinflammation is associated with increased rates of neurodegeneration. Inflammation tends to increase naturally with age. By facilitating inflammation, pollutants such as air particulates and heavy metals cause the CNS to age more quickly. Many late-onset diseases are caused by neurodegeneration. Multiple sclerosis, Parkinson's disease, amyotrophic lateral sclerosis (ALS), and Alzheimer's disease are all believed to be exacerbated by inflammatory processes, resulting in individuals displaying signs of these diseases at an earlier age than is typically expected.

Multiple sclerosis occurs when chronic inflammation leads to the compromise of oligodendrocytes, which in turn leads to the destruction of the myelin sheath. Then axons begin exhibiting signs of damage, which in turn leads to neuron death. Multiple sclerosis has been correlated to living in areas with high particulate matter levels in the air.

In Parkinson's disease, inflammation leading to depletion of antioxidant stores will ultimately lead to dopaminergic neuron degeneration, causing a shortage of dopamine and contributing to the formation of Parkinson's disease. Chronic glial activation as a result of inflammation causes motor neuron death and compromises astrocytes, these factors leading to the symptoms of amyotrophic lateral sclerosis (ALS, aka Lou Gehrig's disease).

In the case of Alzheimer's disease, inflammatory processes lead to neuron death by inhibiting growth at axons and activating astrocytes that produce proteoglycans. This product can only be deposited in the hippocampus and cortex, indicating that this may be the reason these two areas show the highest levels of degeneration in Alzheimer's disease. Airborne metal particulates have been shown to directly access and affect the brain through olfactory pathways, which allows a large amount of particulate matter to reach the blood–brain barrier.

These facts, coupled with air pollution's link to neurofibrillary tangles and the observed subcortical vascular changes observed in dogs, imply that the negative neuroplastic effects of pollution could result in increased risk for Alzheimer's disease, and could also implicate pollution as a cause of early-onset Alzheimer's disease through multiple mechanisms. The general effect of pollution is increased levels of inflammation. As a result, pollution can significantly contribute to various neurological disorders that are caused by inflammatory processes.

World Wide Web Consortium

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/World_Wide_Web_Consortium World Wide We...