Perineuronal net

Last updated
A. Perineuronal nets are made of chondroitin sulfate proteoglycans (CSPGs). Here, the CSPGs neurocan, versican, brevican, and aggrecan are noncovalently bonded to hyaluronan. Associations occur between other CSPGs through tenascin (T, triangles). Tenascin, in turn, binds to CS glycosaminoglycans (red lines) as well as cell surface CSPGs. Phosphacan can also bind to cell surface receptors such as NCAM. B. Application of chondroitinase ABC (ChABC) degrades all the CS glycosaminoglycans (red lines) as well as hyaluronan (pink line), causing major disruptions in the structure of the perineuronal net. These disruptions may allow axons to penetrate the vacated space and permit restoration of neural plasticity. Perineuronal nets and chondroitinase ABC (large).png
A. Perineuronal nets are made of chondroitin sulfate proteoglycans (CSPGs). Here, the CSPGs neurocan, versican, brevican, and aggrecan are noncovalently bonded to hyaluronan. Associations occur between other CSPGs through tenascin (T, triangles). Tenascin, in turn, binds to CS glycosaminoglycans (red lines) as well as cell surface CSPGs. Phosphacan can also bind to cell surface receptors such as NCAM. B. Application of chondroitinase ABC (ChABC) degrades all the CS glycosaminoglycans (red lines) as well as hyaluronan (pink line), causing major disruptions in the structure of the perineuronal net. These disruptions may allow axons to penetrate the vacated space and permit restoration of neural plasticity.

Perineuronal nets (PNNs) are specialized extracellular matrix structures responsible for synaptic stabilization in the adult brain. [1] PNNs are found around certain neuron cell bodies and proximal neurites in the central nervous system. PNNs play a critical role in the closure of the childhood critical period, and their digestion can cause restored critical period-like synaptic plasticity in the adult brain. They are largely negatively charged and composed of chondroitin sulfate proteoglycans, molecules that play a key role in development and plasticity during postnatal development and in the adult.

Contents

PNNs appear to be mainly present in the cortex, hippocampus, thalamus, brainstem, and the spinal cord. Studies of the rat brain have shown that the cortex contains high numbers of PNNs in the motor and primary sensory areas and relatively fewer in the association and limbic cortices. [2] In the cortex, PNNs are associated mostly with inhibitory interneurons and are thought to be responsible for maintaining the excitatory/inhibitory balance in the adult brain. [3]

History

The existence of PNNs has been inferred by Golgi, Lugaro, Donaggio, Martinotti, Ramón y Cajal and Meyer. However, Ramón y Cajal credits Golgi with the discovery of PNNs because he was the first to draw attention to them and gave the first precise description in 1893. Moreover, Golgi brought interest to the subject due to his opinion that the PNN was not a neuronal structure but rather a "kind of corset of neurokeratin which impeded the spread of current from cell to cell". Despite debating the topic, Ramón y Cajal claimed that the perineuronal net was simply a staining artifact derived from the coagulation of extracellular fluids. Due to his influential opinion at the time, interest in the topic subsided.

Interest arose in the 1960s when several authors drew attention to the presence of periodic-acid-Schiff-positive (PAS-positive) material surrounding nerve cells. This PAS-positive material was suspected of being composed of negatively charged substances, such as chondroitin sulfate proteoglycans (CSPGs). However, the authors clung to the idea that the material was intricately connected to the blood–brain barrier and failed to see the similarities it had with the perineuronal net described by Golgi. Interest only rose again in the past few decades when it was discovered that PNNs constitute markers for physiologically mature neurons. [4]

Composition

PNNs are composed of a condensed matrix of chondroitin sulfate proteoglycans, molecules that consist of a core protein and a glycosaminoglycan (GAG) chain. The CS-GAG chains associated with PNNs differs from those found floating in the extra-cellular matrix in a noncondensed form. PNNs are composed of brevican, neurocan, versican, aggrecan, phosphacan, hyaluronan, tenascin-R and various link proteins. The CSPGs aggrecan, versican, neurocan, brevican, and phosphacan are bound to hyaluronan. Many of the components of PNNs are also expressed other forms of ECM in the brain. Aggrecan is selectively expressed in PNNs and is essential for the construction and maintenance of PNNs. [5] PNNs found in both the brain and the spinal cord have the same composition. [6] Chondroitinase ABC (ChABC), a bacterial enzyme routinely used to digest CSPGs, works by catalyzing the removal of the CS-GAG chains of CSPGs, [2] thus it is not selective to PNNs. Mutant mice deficient in tenascin-R or link protein have attenuated PNNs,

In the cortex and other subcortical areas, PNNs preferentially surround GABAergic interneurons containing the calcium-binding protein parvalbumin. [7] [8] The onset of the critical period corresponds closely to the emergence of parvalbumin-positive cells. Parvalbumin-positive cells synapse onto α1-subunit-containing GABAA receptors. The α1-subunit-containing GABAA receptors have been shown to be the only GABAA receptors that drive cortical plasticity. [3] For this reason, PNNs were first thought to have a strong role in the closure of the critical period.

Functions

Neuroprotection

A fine regulation of axonal and dendritic growth is required in the adult CNS to preserve important connections while still allowing for structural plasticity. This function has been recognized to be mediated by several myelin-associated proteins and CSPGs. In order to assess the physiological role of PNNs in the undamaged CNS, ChABC was injected in the healthy cerebellum of adult rats. In the site of ChABC injections, there was profuse outgrowth of the terminal branches of Purkinje cell neurons. However, myelinated axon segments were not affected and remained normal. Purkinje axon sprouting was first evident four days after the degradation of CSPGs. Within 42 days, the expression of CSPGs gradually recovered, at which point axon outgrowth regressed, indicating that there was no significant formation of stable synaptic contacts. While CSPGs are very important in neuroprotection, this indicates that CSPGs may not be the only molecules important for the preservation of anatomical plasticity. [9]

Restriction of AMPA receptor mobility

Cell surface proteins, including neurotransmitter receptors, are highly mobile in the plasma membrane due to lateral diffusion. Fast movements of AMPA-type glutamate receptors (AMPARs) are involved in the modulation of synaptic transmission. As a receptor is used, it becomes desensitized and unable to operate efficiently for a short period of time. Diffusion of the desensitized receptor for the exchange of a naive functional one increases synaptic fidelity during fast repetitive stimulation. PNNs compartmentalize the neuronal surface and act as lateral diffusion barriers for AMPARs, limiting synaptic exchange. This may be part of the reason that synaptic plasticity is limited once PNNs become upregulated. [10]

Buffering system for cations

Most of the parvalbumin-positive neurons surrounded by PNNs also contain the potassium channel Kv3.1b subunit. These specific cells have been identified as fast-spiking cells. These neurons have a low input resistance of the cell membrane, a high resting membrane potential, a short duration of both action potentials and the refractory period, a high firing frequency, and an almost constant amplitude of their action potentials. It appears that both Kv3.1 channels and PNNs are both required for the fast-spiking behavior of these neurons. These potassium channels are important because outward potassium currents are responsible for the repolarization of the cell membrane during an action potential. It has been shown that Kv3.1 currents allow a neuron to follow a high frequency stimulation and/or to generate high firing rates without spike adaption, characteristics that fit well with fast-spiking cells. [11] This characteristic of the cells is important as blockade of the Kv3.1b channel has been shown to slow the rate of ocular dominance plasticity. [3]

PNNs, with their strongly negative charge, may serve as cation exchangers preventing the free diffusion of potassium or sodium ions. Due to the spatial, temporal, and numerical disproportions between Na+ influx and K+ efflux, the PNN provides a possible buffering system for extracellular cations. However, this hypothesis has yet to be proven. [11]

Role in neuroplasticity

PNNs play an important role in neuroplasticity. Traumatic injury of the CNS results in degeneration of denervated and damaged neurons, the formation of a glial scar, and collateral sprouting of surviving neurons. PNNs have been shown to be inhibitory to axonal regeneration and outgrowth. [12] CSPGs are the main axon growth inhibitory molecules in the glial scar that play a role in the failure of the axon to regenerate after injury. [13] In the rat brain and spinal cord, the expression of various CSPGs (brevican, versican, neurocan, and NG2) increases after injury. In vivo treatment with ChABC results in the enhancement of the regeneration of axons (specifically dopaminergic neurons) and the promotion of axon regeneration and functional recovery following spinal cord injury. [2]

CSPGs and PNNs are also implicated in the restricted plasticity present after CNS injury. In the rat cerebellum, application of ChABC promotes structural plasticity of Purkinje axons. [9] Following spinal cord injury, rats treated with ChABC show structural and functional recovery in the form of increased regrowth of axons into the denervated territory and the recovery of motor and bladder function. Plasticity of intact areas in the brain stem and spinal cord also increases following spinal cord injury. [2]

Ocular dominance plasticity

The critical period is a stage when a necessary amount of experience is required for the proper organization of a neural pathway. The absence of this experience may lead to the permanent formation of incorrect connections. [2] The classic model of the critical period has been the visual system. Normally, the primary visual cortex contains neurons organized in ocular dominance columns, with groups of neurons responding preferentially to one eye or the other. If an animal's dominant eye is sutured early in life and kept sutured through the visual critical period (monocular deprivation), the cortex permanently responds preferentially to the eye that was kept open, resulting in ocular dominance shift. However, if the eye is sutured after the critical period, the shift does not occur.

In rats, digestion of PNNs using the bacterial enzyme chondroitinase ABC reactivates the visual critical period. Specifically, digestion of PNNs in the visual cortex well after the closure of the critical period (postnatal day 70) reactivated critical period plasticity and allowed ocular dominance shift to occur. However, the effects of monocular deprivation in the reactivated case were not as strong as monocular deprivation during a normal critical period. [14] Additionally, in adult rats that had been monocularly deprived since youth, digestion of PNNs brought about a full structural and functional recovery (recovery of ocular dominance, visual acuity, and dendritic spine density). However, this recovery only occurred once the open eye was sutured to allow the cortical representation of the deprived eye to recover. [15]

Fear memories

Fear conditioning in animals is used to model anxiety disorders such as PTSD. Fear conditioning works by pairing an initially neutral stimulus with an aversive stimulus, leading to long-lasting fear memories. In an adult animal, fear conditioning induces a permanent memory resilient to erasure by extinction. After extinction, conditioned fear responses can recover spontaneously after a reexposure to the aversive stimulus. In contrast, in early postnatal development, extinction of a conditioned fear response leads to memory erasure. The organization of PNNs in the amygdala coincides with this switch in fear memory resilience. In the adult animal, degradation of PNNs in the amygdala with ChABC renders subsequently acquired fear memories susceptible to erasure. Extinction training was necessary for the loss of fear behavior. Additionally, fear memories acquired before the degradation of the PNNs were not affected by their degradation. [16]

Developmental song learning

Developmental song learning is a model used for the sensorimotor critical period. Birdsong learning in the zebra finch occurs during a critical period similar to that for human speech. This critical period occurs in two parts. The first consists of an early perceptual phase in which sounds are merely memorized. This is followed by a second sensorimotor phase in which feedback is used to shape proper sounds. In the song nuclei HVC, over 80% of PNNs surround parvalbumin-positive neurons. The presence of perineuronal nets predicts the maturity of a zebra finch's song, with greater PNN density indicating a more mature song and likely greater synaptic stability. Unlike the visual critical period, extensive preliminary investigation has shown that degrading the PNNs with ChABC does not reopen the critical period of sensorimotor plasticity. This can be attributed to the additional complicating factors present in a sensorimotor system compared to a purely sensory system. In humans, complications in the speech sensorimotor critical period is implicated in disorders such as autism. Reopening of the critical period in zebra finches may lead to discoveries leading to treatment for these disorders. [17]

Role in CNS pathology

Epilepsy

Epilepsy is a chronic neurological disorder characterized by abnormal electrical activity in the brain. This abnormal electrical activity results in increased plastic changes that play a part of the pathogenesis of the disease. [18] Following seizures, there is a decrease in phosphacan and phosphacan-positive PNNs and an increase in cleaved brevican in the temporal lobe and hippocampus. Seizures also increase the amount of full-length neurocan, a CSPG only found in the neonatal brain. This degradation of CSPGs and PNNs could be responsible for the increased plasticity associated with the disorder. [2]

Stroke

Following stroke, there is some increased plasticity resulting in the restoration of some function. In the rat model, following a cortical lesion, there is a reduction of PNNs in the region surrounding the infarction. Specifically, there is a reduction in the CSPGs aggrecan, versican, and phosphacan and an accumulation of full-length neurocan. This downregulation of PNNs also occurs in brain regions as distant as the thalamus. The degradation of PNNs may be responsible for the increased plasticity seen post-stroke. [19] One issue with typical stroke recovery is the typical period of increased plasticity is generally not long enough to allow stroke patients acceptable recovery of function. One possible treatment strategy may be to degrade PNNs for a longer period of time to allow for greater recovery.

Alzheimer's disease

There appear to be several roles for CSPGs in Alzheimer's disease. PNNs may provide protection against excitotoxicity, oxidative stress, and the formation of neurofibrillary tangles. [2] There have been conflicting reports as to the number of PNNs in the human Alzheimer's brain, with some studies reporting a reduction [20] [21] and others reporting no change. [22] There is no clear consensus as the susceptibility of parvalbumin-positive neurons, the majority of neurons surrounded by PNNs. However, PNNs have been found to localize with both amyloid plaques and neurofibrillary tangles. Since amyloid plaques have been implicated in the progression of Alzheimer's disease, this suggests that PNNs are either instrumental in their formation or are a reaction to their formation. In vitro studies have shown that CSPGs promote beta amyloid fibril formation. Since beta amyloid is a strong stimulant to CSPG production and CSPGs are inhibitory to neuronal growth and synaptic plasticity, this may lead to the decreased axon density and synaptic loss in Alzheimer's disease. [2]

See also

Related Research Articles

<span class="mw-page-title-main">Axon</span> Long projection on a neuron that conducts signals to other neurons

An axon, or nerve fiber, is a long, slender projection of a nerve cell, or neuron, in vertebrates, that typically conducts electrical impulses known as action potentials away from the nerve cell body. The function of the axon is to transmit information to different neurons, muscles, and glands. In certain sensory neurons, such as those for touch and warmth, the axons are called afferent nerve fibers and the electrical impulse travels along these from the periphery to the cell body and from the cell body to the spinal cord along another branch of the same axon. Axon dysfunction can be the cause of many inherited and acquired neurological disorders that affect both the peripheral and central neurons. Nerve fibers are classed into three types – group A nerve fibers, group B nerve fibers, and group C nerve fibers. Groups A and B are myelinated, and group C are unmyelinated. These groups include both sensory fibers and motor fibers. Another classification groups only the sensory fibers as Type I, Type II, Type III, and Type IV.

<span class="mw-page-title-main">Dendrite</span> Small projection on a neuron that receives signals

A dendrite or dendron is a branched protoplasmic extension of a nerve cell that propagates the electrochemical stimulation received from other neural cells to the cell body, or soma, of the neuron from which the dendrites project. Electrical stimulation is transmitted onto dendrites by upstream neurons via synapses which are located at various points throughout the dendritic tree.

<span class="mw-page-title-main">Neuron</span> Electrically excitable cell found in the nervous system of animals

Within a nervous system, a neuron, neurone, or nerve cell is an electrically excitable cell that fires electric signals called action potentials across a neural network. Neurons communicate with other cells via synapses - specialized connections that commonly use minute amounts of chemical neurotransmitters to pass the electric signal from the presynaptic neuron to the target cell through the synaptic gap. The neuron is the main component of nervous tissue in all animals except sponges and placozoa. Non-animals like plants and fungi do not have nerve cells. The ability to generate electric signals first appeared in evolution 700 million years ago. 800 million years ago, predecessors of neurons were the peptidergic secretory cells. They eventually gained new gene modules which enabled cells to create post-synaptic scaffolds and ion channels that generate fast electrical signals. The ability to generate electric signals was a key innovation in the evolution of the nervous system.

<span class="mw-page-title-main">Chemical synapse</span> Biological junctions through which neurons signals can be sent

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.

The development of the nervous system, or neural development (neurodevelopment), refers to the processes that generate, shape, and reshape the nervous system of animals, from the earliest stages of embryonic development to adulthood. The field of neural development draws on both neuroscience and developmental biology to describe and provide insight into the cellular and molecular mechanisms by which complex nervous systems develop, from nematodes and fruit flies to mammals.

<span class="mw-page-title-main">Pyramidal cell</span> Projection neurons in the cerebral cortex and hippocampus

Pyramidal cells, or pyramidal neurons, are a type of multipolar neuron found in areas of the brain including the cerebral cortex, the hippocampus, and the amygdala. Pyramidal cells are the primary excitation units of the mammalian prefrontal cortex and the corticospinal tract. Pyramidal neurons are also one of two cell types where the characteristic sign, Negri bodies, are found in post-mortem rabies infection. Pyramidal neurons were first discovered and studied by Santiago Ramón y Cajal. Since then, studies on pyramidal neurons have focused on topics ranging from neuroplasticity to cognition.

Synaptogenesis is the formation of synapses between neurons in the nervous system. Although it occurs throughout a healthy person's lifespan, an explosion of synapse formation occurs during early brain development, known as exuberant synaptogenesis. Synaptogenesis is particularly important during an individual's critical period, during which there is a certain degree of synaptic pruning due to competition for neural growth factors by neurons and synapses. Processes that are not used, or inhibited during their critical period will fail to develop normally later on in life.

<span class="mw-page-title-main">Barrel cortex</span> Region of the somatosensory cortex in some rodents and other species

The barrel cortex is a region of the somatosensory cortex that is identifiable in some species of rodents and species of at least two other orders and contains the barrel field. The 'barrels' of the barrel field are regions within cortical layer IV that are visibly darker when stained to reveal the presence of cytochrome c oxidase and are separated from each other by lighter areas called septa. These dark-staining regions are a major target for somatosensory inputs from the thalamus, and each barrel corresponds to a region of the body. Due to this distinctive cellular structure, organisation, and functional significance, the barrel cortex is a useful tool to understand cortical processing and has played an important role in neuroscience. The majority of what is known about corticothalamic processing comes from studying the barrel cortex, and researchers have intensively studied the barrel cortex as a model of neocortical column.

In developmental psychology and developmental biology, a critical period is a maturational stage in the lifespan of an organism during which the nervous system is especially sensitive to certain environmental stimuli. If, for some reason, the organism does not receive the appropriate stimulus during this "critical period" to learn a given skill or trait, it may be difficult, ultimately less successful, or even impossible, to develop certain associated functions later in life. Functions that are indispensable to an organism's survival, such as vision, are particularly likely to develop during critical periods. "Critical period" also relates to the ability to acquire one's first language. Researchers found that people who passed the "critical period" would not acquire their first language fluently.

Ocular dominance columns are stripes of neurons in the visual cortex of certain mammals that respond preferentially to input from one eye or the other. The columns span multiple cortical layers, and are laid out in a striped pattern across the surface of the striate cortex (V1). The stripes lie perpendicular to the orientation columns.

<span class="mw-page-title-main">Aggrecan</span>

Aggrecan (ACAN), also known as cartilage-specific proteoglycan core protein (CSPCP) or chondroitin sulfate proteoglycan 1, is a protein that in humans is encoded by the ACAN gene. This gene is a member of the lectican (chondroitin sulfate proteoglycan) family. The encoded protein is an integral part of the extracellular matrix in cartilagenous tissue and it withstands compression in cartilage.

<span class="mw-page-title-main">Chandelier cell</span>

Chandelier neurons or chandelier cells are a subset of GABAergic cortical interneurons. They are described as parvalbumin-containing and fast-spiking to distinguish them from other subtypes of GABAergic neurons, although more recent work has suggested that only a subset of chandelier cells test positive for parvalbumin by immunostaining. The name comes from the specific shape of their axon arbors, with the terminals forming distinct arrays called "cartridges". The cartridges are immunoreactive to an isoform of the GABA membrane transporter, GAT-1, and this serves as their identifying feature. GAT-1 is involved in the process of GABA reuptake into nerve terminals, thus helping to terminate its synaptic activity. Chandelier neurons synapse exclusively to the axon initial segment of pyramidal neurons, near the site where action potential is generated. It is believed that they provide inhibitory input to the pyramidal neurons, but there is data showing that in some circumstances the GABA from chandelier neurons could be excitatory.

Neuroregeneration involves the regrowth or repair of nervous tissues, cells or cell products. Neuroregenerative mechanisms may include generation of new neurons, glia, axons, myelin, or synapses. Neuroregeneration differs between the peripheral nervous system (PNS) and the central nervous system (CNS) by the functional mechanisms involved, especially in the extent and speed of repair. When an axon is damaged, the distal segment undergoes Wallerian degeneration, losing its myelin sheath. The proximal segment can either die by apoptosis or undergo the chromatolytic reaction, which is an attempt at repair. In the CNS, synaptic stripping occurs as glial foot processes invade the dead synapse.

Chondroitinase treatment is a treatment of proteoglycans, a protein in the fluid among cells where they affect neural activity. Chondroitinase treatment has been shown to allow adults vision to be restored as far as ocular dominance is concerned. Moreover, there is some evidence that Chondroitinase could be used for the treatment of spinal injuries.

<span class="mw-page-title-main">Glial scar</span> Mass formed in response to injury to the nervous system

A glial scar formation (gliosis) is a reactive cellular process involving astrogliosis that occurs after injury to the central nervous system. As with scarring in other organs and tissues, the glial scar is the body's mechanism to protect and begin the healing process in the nervous system.

<span class="mw-page-title-main">ADAMTS4</span> Protein-coding gene in the species Homo sapiens

A disintegrin and metalloproteinase with thrombospondin motifs 4 is an enzyme that in humans is encoded by the ADAMTS4 gene.

<span class="mw-page-title-main">Nonsynaptic plasticity</span> Form of neuroplasticity

Nonsynaptic plasticity is a form of neuroplasticity that involves modification of ion channel function in the axon, dendrites, and cell body that results in specific changes in the integration of excitatory postsynaptic potentials and inhibitory postsynaptic potentials. Nonsynaptic plasticity is a modification of the intrinsic excitability of the neuron. It interacts with synaptic plasticity, but it is considered a separate entity from synaptic plasticity. Intrinsic modification of the electrical properties of neurons plays a role in many aspects of plasticity from homeostatic plasticity to learning and memory itself. Nonsynaptic plasticity affects synaptic integration, subthreshold propagation, spike generation, and other fundamental mechanisms of neurons at the cellular level. These individual neuronal alterations can result in changes in higher brain function, especially learning and memory. However, as an emerging field in neuroscience, much of the knowledge about nonsynaptic plasticity is uncertain and still requires further investigation to better define its role in brain function and behavior.

<span class="mw-page-title-main">Chondroitin sulfate proteoglycan</span>

Chondroitin sulfate proteoglycans (CSPGs) are proteoglycans consisting of a protein core and a chondroitin sulfate side chain. They are known to be structural components of a variety of human tissues, including cartilage, and also play key roles in neural development and glial scar formation. They are known to be involved in certain cell processes, such as cell adhesion, cell growth, receptor binding, cell migration, and interaction with other extracellular matrix constituents. They are also known to interact with laminin, fibronectin, tenascin, and collagen. CSPGs are generally secreted from cells.

In neuroscience, synaptic scaling is a form of homeostatic plasticity, in which the brain responds to chronically elevated activity in a neural circuit with negative feedback, allowing individual neurons to reduce their overall action potential firing rate. Where Hebbian plasticity mechanisms modify neural synaptic connections selectively, synaptic scaling normalizes all neural synaptic connections by decreasing the strength of each synapse by the same factor, so that the relative synaptic weighting of each synapse is preserved.

<span class="mw-page-title-main">Cortical remapping</span>

Cortical remapping, also referred to as cortical reorganization, is the process by which an existing cortical map is affected by a stimulus resulting in the creating of a 'new' cortical map. Every part of the body is connected to a corresponding area in the brain which creates a cortical map. When something happens to disrupt the cortical maps such as an amputation or a change in neuronal characteristics, the map is no longer relevant. The part of the brain that is in charge of the amputated limb or neuronal change will be dominated by adjacent cortical regions that are still receiving input, thus creating a remapped area. Remapping can occur in the sensory or motor system. The mechanism for each system may be quite different. Cortical remapping in the somatosensory system happens when there has been a decrease in sensory input to the brain due to deafferentation or amputation, as well as a sensory input increase to an area of the brain. Motor system remapping receives more limited feedback that can be difficult to interpret.

References

  1. Flores, CE; Méndez, P (2014). "Shaping inhibition: activity dependent structural plasticity of GABAergic synapses". Frontiers in Cellular Neuroscience. 8: 327. doi: 10.3389/fncel.2014.00327 . PMC   4209871 . PMID   25386117.
  2. 1 2 3 4 5 6 7 8 Galtrey, C. M.; Fawcett, J. W. (2007). "The role of chondroitin sulfate proteoglycans in regeneration and plasticity in the central nervous system". Brain Research Reviews. 54 (1): 1–18. doi:10.1016/j.brainresrev.2006.09.006. PMID   17222456. S2CID   13913614.
  3. 1 2 3 Hensch, T. K. (2005). "Critical period plasticity in local cortical circuits. [Review]". Nature Reviews Neuroscience. 6 (11): 877–888. doi:10.1038/nrn1787. hdl: 10533/174307 . PMID   16261181. S2CID   5264124.
  4. Celio, M. R.; Spreafico, R.; De Biasi, S.; Vitellaro-Zuccarello, L. (1998). "Perineuronal nets: past and present". Trends in Neurosciences. 21 (12): 510–515. doi:10.1016/s0166-2236(98)01298-3. PMID   9881847. S2CID   28557831.
  5. Rowlands, Daire; Lensjø, Kristian K.; Dinh, Tovy; Yang, Sujeong; Andrews, Melissa R.; Hafting, Torkel; Fyhn, Marianne; Fawcett, James W.; Dick, Gunnar (2018-11-21). "Aggrecan Directs Extracellular Matrix-Mediated Neuronal Plasticity". Journal of Neuroscience. 38 (47): 10102–10113. doi: 10.1523/JNEUROSCI.1122-18.2018 . ISSN   0270-6474. PMC   6596198 . PMID   30282728.
  6. Deepa, S. S.; Carulli, D.; Galtrey, C.; Rhodes, K.; Fukuda, J.; Mikami, T.; Fawcett, J. W. (2006). "Composition of Perineuronal Net Extracellular Matrix in Rat Brain". Journal of Biological Chemistry. 281 (26): 17789–17800. doi: 10.1074/jbc.m600544200 . PMID   16644727.
  7. Morris, N.P.; Henderson, Z. (2000). "Perineuronal nets ensheath fast spiking, parvalbumin-immunoreactive neurons in the medial septum/diagonal band complex". Eur J Neurosci. 12 (3): 828–38. doi:10.1046/j.1460-9568.2000.00970.x. PMID   10762312. S2CID   22232630.
  8. Mayne, PE; Burne, THJ (April 2019). "Vitamin D in Synaptic Plasticity, Cognitive Function, and Neuropsychiatric Illness". Trends in Neurosciences. 42 (4): 293–306. doi:10.1016/j.tins.2019.01.003. PMID   30795846. S2CID   73479985.
  9. 1 2 Corvetti, L.; Rossi, F. (2005). "Degradation of Chondroitin Sulfate Proteoglycans Induces Sprouting of Intact Purkinje Axons in the Cerebellum of the Adult Rat". J. Neurosci. 25 (31): 7150–7158. doi:10.1523/jneurosci.0683-05.2005. PMC   6725229 . PMID   16079397.
  10. Frischknecht, R.; Heine, M.; Perrais, D.; Seidenbecher, C. I.; Choquet, D.; Gundelfinger, E. D. (2009). "Brain extracellular matrix affects AMPA receptor lateral mobility and short-term synaptic plasticity. [10.1038/nn.2338]". Nat Neurosci. 12 (7): 897–904. doi:10.1038/nn.2338. PMID   19483686. S2CID   1005267.
  11. 1 2 Hartig, W.; Derouiche, A.; Welt, K.; Brauer, K.; Grosche, J.; Mader, M.; et al. (1999). "Cortical neurons immunoreactive for the potassium channel Kv3.1b subunit are predominantly surrounded by perineuronal nets presumed as a buffering system for cations. [Article]". Brain Research. 842 (1): 15–29. doi:10.1016/s0006-8993(99)01784-9. PMID   10526091. S2CID   19980614.
  12. Rhodes, K. E.; Fawcett, J. W. (2004). "Chondroitin sulphate proteoglycans: preventing plasticity or protecting the CNS? [Review]". Journal of Anatomy. 204 (1): 33–48. doi:10.1111/j.1469-7580.2004.00261.x. PMC   1571240 . PMID   14690476.
  13. Silver, J.; Miller, J. H. (2004). "Regeneration beyond the glial scar. [10.1038/nrn1326]". Nat Rev Neurosci. 5 (2): 146–156. doi: 10.1038/nrn1326 . PMID   14735117.
  14. Pizzorusso, T.; Medini, P.; Berardi, N.; Chierzi, S.; Fawcett, J. W.; Maffei, L. (2002). "Reactivation of ocular dominance plasticity in the adult visual cortex. [Article]". Science. 298 (5596): 1248–1251. Bibcode:2002Sci...298.1248P. doi:10.1126/science.1072699. PMID   12424383. S2CID   14254863.
  15. Pizzorusso, T.; Medini, P.; Landi, S.; Baldini, S.; Berardi, N.; Maffei, L. (2006). "Structural and functional recovery from early monocular deprivation in adult rats". Proceedings of the National Academy of Sciences. 103 (22): 8517–8522. Bibcode:2006PNAS..103.8517P. doi: 10.1073/pnas.0602657103 . PMC   1482523 . PMID   16709670.
  16. Gogolla, N.; Caroni, P.; Lüthi, A.; Herry, C. (2009). "Perineuronal Nets Protect Fear Memories from Erasure". Science. 325 (5945): 1258–1261. Bibcode:2009Sci...325.1258G. doi:10.1126/science.1174146. PMID   19729657. S2CID   206520056.
  17. Balmer, T. S.; Carels, V. M.; Frisch, J. L.; Nick, T. A. (2009). "Modulation of Perineuronal Nets and Parvalbumin with Developmental Song Learning". Journal of Neuroscience. 29 (41): 12878–12885. doi:10.1523/jneurosci.2974-09.2009. PMC   2769505 . PMID   19828802.
  18. Morimoto, K.; Fahnestock, M.; Racine, R. J. (2004). "Kindling and status epilepticus models of epilepsy: rewiring the brain". Progress in Neurobiology. 73 (1): 1–60. doi:10.1016/j.pneurobio.2004.03.009. PMID   15193778. S2CID   36849482.
  19. Bidmon, H. J.; Jancsik, V.; Schleicher, A.; Hagemann, G.; Witte, O. W.; Woodhams, P.; Zilles, K. (1997). "Structural alterations and changes in cytoskeletal proteins and proteoglycans after focal cortical ischemia". Neuroscience. 82 (2): 397–420. doi:10.1016/s0306-4522(97)00289-3. PMID   9466450. S2CID   42361088.
  20. Kobayashi, K.; Emson, P. C.; Mountjoy, C. Q. (1989). "Vicia villosa lectin-positive neurones in human cerebral cortex. Loss in Alzheimer-type dementia". Brain Research. 498 (1): 170–174. doi:10.1016/0006-8993(89)90416-2. PMID   2790470. S2CID   37981595.
  21. Baig, S.; Wilcock, G.; Love, S. (2005). "Loss of perineuronal net<i> N</i>-acetylgalactosamine in Alzheimer's disease". Acta Neuropathologica. 110 (4): 393–401. doi:10.1007/s00401-005-1060-2. PMID   16133543. S2CID   27218812.
  22. Brückner, G.; Hausen, D.; Härtig, W.; Drlicek, M.; Arendt, T.; Brauer, K. (1999). "Cortical areas abundant in extracellular matrix chondroitin sulphate proteoglycans are less affected by cytoskeletal changes in Alzheimer's disease". Neuroscience. 92 (3): 791–805. doi:10.1016/s0306-4522(99)00071-8. PMID   10426522. S2CID   24281926.