Voltage-gated sodium channel

Last updated

Voltage-gated sodium channels (VGSCs), also known as voltage-dependent sodium channels (VDSCs), are a group of voltage-gated ion channels found in the membrane of excitable cells (e.g., muscle, glial cells, neurons, etc.) with a permeability to the sodium ion Na+. They are the main channels involved in action potential of excitable cells.

Contents

Structure

Diagram of a voltage-sensitive sodium channel a-subunit. G - glycosylation, P - phosphorylation, S - ion selectivity, I - inactivation. Positive (+) charges in S4 are important for transmembrane voltage sensing. Sodium-channel.svg
Diagram of a voltage-sensitive sodium channel α-subunit. G – glycosylation, P – phosphorylation, S – ion selectivity, I – inactivation. Positive (+) charges in S4 are important for transmembrane voltage sensing.

Sodium channels consist of large alpha subunits that associate with accessory proteins, such as beta subunits. An alpha subunit forms the core of the channel and is functional on its own. When the alpha subunit protein is expressed by a cell, it is able to form a pore in the cell membrane that conducts Na+ in a voltage-dependent way, even if beta subunits or other known modulating proteins are not expressed. When accessory proteins assemble with α subunits, the resulting complex can display altered voltage dependence and cellular localization.

The alpha subunit consists of four repeat domains, labelled I through IV, each containing six membrane-spanning segments, labelled S1 through S6. The highly conserved S4 segment acts as the channel's voltage sensor. The voltage sensitivity of this channel is due to positive amino acids located at every third position. [2] When stimulated by a change in transmembrane voltage, this segment moves toward the extracellular side of the cell membrane, allowing the channel to become permeable to ions. The ions are conducted through the central pore cavity, which consists of two main regions. The more external (i.e., more extracellular) portion of the pore is formed by the "P-loops" (the region between S5 and S6) of the four domains. This region is the most narrow part of the pore and is responsible for its ion selectivity. The inner portion (i.e., more cytoplasmic) of the pore is the pore gate and is formed by the combined S5 and S6 segments of the four domains. The pore domain also features lateral tunnels or fenestrations that run perpendicular to the pore axis. These fenestrations that connect the central cavity to the membrane are proposed to be important for drug accessibility. [3] [4] [5]

In mammalian sodium channels, the region linking domains III and IV is also important for channel function. This DIII-IV linker is responsible for wedging the pore gate shut after channel opening, inactivating it. [6]

Gating

Voltage-gated Na+ channels have three main conformational states: closed, open and inactivated. Forward/back transitions between these states are correspondingly referred to as activation/deactivation (between open and closed, respectively), inactivation/reactivation (between inactivated and open, respectively), and recovery from inactivation/closed-state inactivation (between inactivated and closed, respectively). Closed and inactivated states are ion impermeable.

Before an action potential occurs, the axonal membrane is at its normal resting potential, about −70 mV in most human neurons, and Na+ channels are in their deactivated state, blocked on the extracellular side by their activation gates. In response to an increase of the membrane potential to about −55 mV (in this case, caused by an action potential), the activation gates open, allowing positively charged Na+ ions to flow into the neuron through the channels, and causing the voltage across the neuronal membrane to increase to +30 mV in human neurons. Because the voltage across the membrane is initially negative, as its voltage increases to and past zero (from −70 mV at rest to a maximum of +30 mV), it is said to depolarize. This increase in voltage constitutes the rising phase of an action potential.

Action PotentialMembrane PotentialTarget PotentialGate's Target StateNeuron's Target State
Resting−70 mV−55 mVDeactivated → ActivatedPolarized
Rising−55 mV0 mVActivatedPolarized → Depolarized
Rising0 mV+30 mVActivated → InactivatedDepolarized
Falling+30 mV0 mVInactivatedDepolarized → Repolarized
Falling0 mV−70 mVInactivatedRepolarized
Undershot−70 mV−75 mVInactivated → DeactivatedRepolarized → Hyperpolarized
Rebounding−75 mV−70 mVDeactivatedHyperpolarized → Polarized

At the peak of the action potential, when enough Na+ has entered the neuron and the membrane's potential has become high enough, the Na+ channels inactivate themselves by closing their inactivation gates . The inactivation gate can be thought of as a "plug" tethered to domains III and IV of the channel's intracellular alpha subunit. Closure of the inactivation gate causes Na+ flow through the channel to stop, which in turn causes the membrane potential to stop rising. The closing of the inactivation gate creates a refractory period within each individual Na+ channel. This refractory period eliminates the possibility of an action potential moving in the opposite direction back towards the soma. With its inactivation gate closed, the channel is said to be inactivated. With the Na+ channel no longer contributing to the membrane potential, the potential decreases back to its resting potential as the neuron repolarizes and subsequently hyperpolarizes itself, and this constitutes the falling phase of an action potential. The refractory period of each channel is therefore vital in propagating the action potential unidirectionally down an axon for proper communication between neurons.

When the membrane's voltage becomes low enough, the inactivation gate reopens and the activation gate closes in a process called deinactivation. With the activation gate closed and the inactivation gate open, the Na+ channel is once again in its deactivated state, and is ready to participate in another action potential.

When any kind of ion channel does not inactivate itself, it is said to be persistently (or tonically) active. Some kinds of ion channels are naturally persistently active. However, genetic mutations that cause persistent activity in other channels can cause disease by creating excessive activity of certain kinds of neurons. Mutations that interfere with Na+ channel inactivation can contribute to cardiovascular diseases or epileptic seizures by window currents, which can cause muscle and/or nerve cells to become over-excited.

Modeling the behavior of gates

The temporal behavior of Na+ channels can be modeled by a Markovian scheme or by the Hodgkin–Huxley-type formalism. In the former scheme, each channel occupies a distinct state with differential equations describing transitions between states; in the latter, the channels are treated as a population that are affected by three independent gating variables. Each of these variables can attain a value between 1 (fully permeant to ions) and 0 (fully non-permeant), the product of these variables yielding the percentage of conducting channels. The Hodgkin–Huxley model can be shown to be equivalent to a Markovian model.[ further explanation needed ]

Impermeability to other ions

The pore of sodium channels contains a selectivity filter made of negatively charged amino acid residues, which attract the positive Na+ ion and keep out negatively charged ions such as chloride. The cations flow into a more constricted part of the pore that is 0.3 by 0.5 nm wide, which is just large enough to allow a single Na+ ion with a water molecule associated to pass through. The larger K+ ion cannot fit through this area. Ions of different sizes also cannot interact as well with the negatively charged glutamic acid residues that line the pore. [ citation needed ]

Diversity

Voltage-gated sodium channels normally consist of an alpha subunit that forms the ion conduction pore and one to two beta subunits that have several functions including modulation of channel gating. [7] Expression of the alpha subunit alone is sufficient to produce a functional channel.

Alpha subunits

Figure 1. Likely evolutionary relationship of the nine known human sodium channels. Sodium channel phylogram.png
Figure 1. Likely evolutionary relationship of the nine known human sodium channels.

The family of sodium channels has 9 known members, with amino acid identity >50% in the trans-membrane segments and extracellular loop regions. A standardized nomenclature for sodium channels is currently used and is maintained by the IUPHAR. [8]

The proteins of these channels are named Nav1.1 through Nav1.9. The gene names are referred to as SCN1A through SCN5A, then SCN8A through SCN11A. [8] The "tenth member", Nax, does not act in a voltage-gated way. It has a loosely similar overall structure. Not much is known about its real function, other than that it also associates with beta subunits. [9]

The probable evolutionary relationship between these channels, based on the similarity of their amino acid sequences, is shown in figure 1. The individual sodium channels are distinguished not only by differences in their sequence but also by their kinetics and expression profiles. Some of this data is summarized in table 1, below.

Table 1. Nomenclature and some functions of voltage-gated sodium channel alpha subunits
Protein nameGeneExpression profileAssociated human channelopathies
Nav1.1 SCN1A Central neurons, [peripheral neurons] and cardiac myocytes febrile epilepsy, GEFS+, Dravet syndrome (also known as severe myclonic epilepsy of infancy or SMEI), borderline SMEI (SMEB), West syndrome (also known as infantile spasms), Doose syndrome (also known as myoclonic astatic epilepsy), intractable childhood epilepsy with generalized tonic-clonic seizures (ICEGTC), Panayiotopoulos syndrome, familial hemiplegic migraine (FHM), familial autism, Rasmussens's encephalitis and Lennox-Gastaut syndrome [10]
Nav1.2 SCN2A Central neurons, peripheral neuronsinherited febrile seizures, epilepsy, and autism spectrum disorder
Nav1.3 SCN3A Central neurons, peripheral neurons and cardiac myocytesepilepsy, pain, brain malformations [11] [12]
Nav1.4 SCN4A Skeletal muscle hyperkalemic periodic paralysis, paramyotonia congenita, and potassium-aggravated myotonia
Nav1.5 SCN5A Cardiac myocytes, uninnervated skeletal muscle, central neurons, gastrointestinal smooth muscle cells and Interstitial cells of CajalCardiac: Long QT syndrome Type 3, Brugada syndrome, progressive cardiac conduction disease, familial atrial fibrillation and idiopathic ventricular fibrillation; [13]

Gastrointestinal: Irritable bowel syndrome; [14]

Nav1.6 SCN8A Central neurons, dorsal root ganglia, peripheral neurons, heart, glia cells Epilepsy, [15] ataxia, dystonia, tremor [16]
Nav1.7 SCN9A Dorsal root ganglia, sympathetic neurons, Schwann cells, and neuroendocrine cells erythromelalgia, PEPD, channelopathy-associated insensitivity to pain [11] and recently discovered a disabling form of fibromyalgia (rs6754031 polymorphism) [17]
Nav1.8 SCN10A Dorsal root gangliapain, [11] neuropsychiatric disorders
Nav1.9 SCN11A Dorsal root gangliapain [11]
Nax SCN7A heart, uterus, skeletal muscle, astrocytes, dorsal root ganglion cellsnone known

Beta subunits

Sodium channel beta subunits are type 1 transmembrane glycoproteins with an extracellular N-terminus and a cytoplasmic C-terminus. As members of the Ig superfamily, beta subunits contain a prototypic V-set Ig loop in their extracellular domain. They do not share any homology with their counterparts of calcium and potassium channels. [18] Instead, they are homologous to neural cell adhesion molecules (CAMs) and the large family of L1 CAMs. There are four distinct betas named in order of discovery: SCN1B, SCN2B, SCN3B, SCN4B (table 2). Beta 1 and beta 3 interact with the alpha subunit non-covalently, whereas beta 2 and beta 4 associate with alpha via disulfide bond. [19] Sodium channels are more likely to stay open at the subthreshold membrane potential when interacting with beta toxins, which in turn induces an immediate sensation of pain. [20]

Role of beta subunits as cell adhesion molecules

In addition to regulating channel gating, sodium channel beta subunits also modulate channel expression and form links to the intracellular cytoskeleton via ankyrin and spectrin. [7] [21] [22] Voltage-gated sodium channels also assemble with a variety of other proteins, such as FHF proteins (Fibroblast growth factor Homologous Factor), calmodulin, cytoskeleton or regulatory kinases, [23] [7] [24] [25] [26] which form a complex with sodium channels, influencing its expression and/or function. Several beta subunits interact with one or more extracellular matrix (ECM) molecules. Contactin, also known as F3 or F11, associates with beta 1 as shown via co-immunoprecipitation. [27] Fibronectin-like (FN-like) repeats of Tenascin-C and Tenascin-R bind with beta 2 in contrast to the Epidermal growth factor-like (EGF-like) repeats that repel beta2. [28] A disintegrin and metalloproteinase (ADAM) 10 sheds beta 2's ectodomain possibly inducing neurite outgrowth. [29] Beta 3 and beta 1 bind to neurofascin at Nodes of Ranvier in developing neurons. [30]

Table 2. Nomenclature and some functions of voltage-gated sodium channel beta subunits
Protein nameGene linkAssembles withExpression profileAssociated human channelopathies
Navβ1 SCN1B Nav1.1 to Nav1.7Central Neurons, Peripheral Neurons, skeletal muscle, heart, gliaepilepsy (GEFS+), Brugada syndrome [31]
Navβ2 SCN2B Nav1.1, Nav1.2, Nav1.5 to Nav1.7Central Neurons, peripheral neurons, heart, glia Brugada syndrome [31]
Navβ3 SCN3B Nav1.1 to Nav1.3, Nav1.5central neurons, adrenal gland, kidney, peripheral neurons Brugada syndrome [31]
Navβ4 SCN4B Nav1.1, Nav1.2, Nav1.5heart, skeletal muscle, central and peripheral neuronsnone known

Evolution

A voltage-gated sodium channel is present in members of the choanoflagellates, thought to be the closest living, unicellular relative of animals. [32] [33] This suggests that an ancestral form of the animal channel was among the many proteins that play central roles in animal life, but which are thought to have evolved before multicellularity. [34] The four-domain animal voltage-gated sodium channel likely evolved from a single-subunit ion channel, which was probably permeable for potassium ions, via a sequence of two duplication events. [35] This model draws support from the fact that subunits I and III (and II and IV) group by similarity, suggesting that a two-channel intermediate generated from the first duplication existed long enough for divergence to occur between its two subunits. After the second duplication, the channel was left with two sets of similar domains. [35] The resulting four-domain channel is thought to have been permeable primarily for calcium, and to have achieved sodium selectivity a number of times independently. [36] [37] After divergence from the invertebrates, the vertebrate lineage underwent two whole-genome duplications (WGDs), yielding a set of four sodium channel gene prologues in the ancestral vertebrate, all of which were retained. [38] [39] After the tetrapod/teleost split, the teleosts likely underwent a third WGD leading to the eight sodium channel prologues expressed in many modern fishes. [38] The modern, ten-paralogue sodium gene complement of mammals is thought to have arisen from a series of parallel and nested duplications involving two of the four paralogues present in the ancestor of all tetrapods. [39]

In electric fishes

Several fishes possess an electric organ functioning with voltage-gated sodium channels, allowing them to communicate, immobilize preys or deter predators, depending on the fish. Electric organs used in communication have evolved independently at least twice: in the Gymnotiformes of South America and the Mormyriformes of Africa of Africa. [40] In non-electric fishes the channels Nav1.4a and Nav1.4b are present in the muscles, while in Gymnotiformes and Mormyriformes Nav1.4a has disapeared in the muscles and is expressed in the electric organ. [40]

Sources

Creative Commons by small.svg  This article incorporates text by Maryam Iman, Atefeh Saadabadi, and Asghar Davood available under the CC BY 4.0 license.

Related Research Articles

Hyperkalemic periodic paralysis is an inherited autosomal dominant disorder that affects sodium channels in muscle cells and the ability to regulate potassium levels in the blood. It is characterized by muscle hyperexcitability or weakness which, exacerbated by potassium, heat or cold, can lead to uncontrolled shaking followed by paralysis. Onset usually occurs in early childhood, but it still occurs with adults.

<span class="mw-page-title-main">Voltage-gated ion channel</span> Type of ion channel transmembrane protein

Voltage-gated ion channels are a class of transmembrane proteins that form ion channels that are activated by changes in the electrical membrane potential near the channel. The membrane potential alters the conformation of the channel proteins, regulating their opening and closing. Cell membranes are generally impermeable to ions, thus they must diffuse through the membrane through transmembrane protein channels. They have a crucial role in excitable cells such as neuronal and muscle tissues, allowing a rapid and co-ordinated depolarization in response to triggering voltage change. Found along the axon and at the synapse, voltage-gated ion channels directionally propagate electrical signals. Voltage-gated ion-channels are usually ion-specific, and channels specific to sodium (Na+), potassium (K+), calcium (Ca2+), and chloride (Cl) ions have been identified. The opening and closing of the channels are triggered by changing ion concentration, and hence charge gradient, between the sides of the cell membrane.

Voltage-gated calcium channels (VGCCs), also known as voltage-dependent calcium channels (VDCCs), are a group of voltage-gated ion channels found in the membrane of excitable cells (e.g., muscle, glial cells, neurons, etc.) with a permeability to the calcium ion Ca2+. These channels are slightly permeable to sodium ions, so they are also called Ca2+–Na+ channels, but their permeability to calcium is about 1000-fold greater than to sodium under normal physiological conditions.

Sodium channels are integral membrane proteins that form ion channels, conducting sodium ions (Na+) through a cell's membrane. They belong to the superfamily of cation channels.

<span class="mw-page-title-main">Hodgkin–Huxley model</span> Describes how neurons transmit electric signals

The Hodgkin–Huxley model, or conductance-based model, is a mathematical model that describes how action potentials in neurons are initiated and propagated. It is a set of nonlinear differential equations that approximates the electrical engineering characteristics of excitable cells such as neurons and muscle cells. It is a continuous-time dynamical system.

<span class="mw-page-title-main">Voltage-gated potassium channel</span> Class of transport proteins

Voltage-gated potassium channels (VGKCs) are transmembrane channels specific for potassium and sensitive to voltage changes in the cell's membrane potential. During action potentials, they play a crucial role in returning the depolarized cell to a resting state.

Na<sub>v</sub>1.4 Protein-coding gene in the species Homo sapiens

Sodium channel protein type 4 subunit alpha is a protein that in humans is encoded by the SCN4A gene.

SCN5A Protein-coding gene in the species Homo sapiens

Sodium channel protein type 5 subunit alpha, also known as NaV1.5 is an integral membrane protein and tetrodotoxin-resistant voltage-gated sodium channel subunit. NaV1.5 is found primarily in cardiac muscle, where it mediates the fast influx of Na+-ions (INa) across the cell membrane, resulting in the fast depolarization phase of the cardiac action potential. As such, it plays a major role in impulse propagation through the heart. A vast number of cardiac diseases is associated with mutations in NaV1.5 (see paragraph genetics). SCN5A is the gene that encodes the cardiac sodium channel NaV1.5.

Na<sub>v</sub>1.7 Protein-coding gene in the species Homo sapiens

Nav1.7 is a sodium ion channel that in humans is encoded by the SCN9A gene. It is usually expressed at high levels in two types of neurons: the nociceptive (pain) neurons at dorsal root ganglion (DRG) and trigeminal ganglion and sympathetic ganglion neurons, which are part of the autonomic (involuntary) nervous system.

T-type calcium channels are low voltage activated calcium channels that become inactivated during cell membrane hyperpolarization but then open to depolarization. The entry of calcium into various cells has many different physiological responses associated with it. Within cardiac muscle cell and smooth muscle cells voltage-gated calcium channel activation initiates contraction directly by allowing the cytosolic concentration to increase. Not only are T-type calcium channels known to be present within cardiac and smooth muscle, but they also are present in many neuronal cells within the central nervous system. Different experimental studies within the 1970s allowed for the distinction of T-type calcium channels from the already well-known L-type calcium channels. The new T-type channels were much different from the L-type calcium channels due to their ability to be activated by more negative membrane potentials, had small single channel conductance, and also were unresponsive to calcium antagonist drugs that were present. These distinct calcium channels are generally located within the brain, peripheral nervous system, heart, smooth muscle, bone, and endocrine system.

Na<sub>v</sub>1.9 Protein-coding gene in the species Homo sapiens

Sodium channel, voltage-gated, type XI, alpha subunit also known as SCN11A or Nav1.9 is a voltage-gated sodium ion channel protein which is encoded by the SCN11A gene on chromosome 3 in humans. Like Nav1.7 and Nav1.8, Nav1.9 plays a role in pain perception. This channel is largely expressed in small-diameter nociceptors of the dorsal root ganglion and trigeminal ganglion neurons, but is also found in intrinsic myenteric neurons.

SCN1A Protein-coding gene in the species Homo sapiens

Sodium channel protein type 1 subunit alpha (SCN1A), is a protein which in humans is encoded by the SCN1A gene.

<span class="mw-page-title-main">SCN1B</span> Protein-coding gene in the species Homo sapiens

Sodium channel subunit beta-1 is a protein that in humans is encoded by the SCN1B gene.

<span class="mw-page-title-main">SCN8A</span> Protein-coding gene in the species Homo sapiens

Sodium channel protein type 8 subunit alpha also known as Nav1.6 is a membrane protein encoded by the SCN8A gene. Nav1.6 is one sodium channel isoform and is the primary voltage-gated sodium channel at each node of Ranvier. The channels are highly concentrated in sensory and motor axons in the peripheral nervous system and cluster at the nodes in the central nervous system.

<span class="mw-page-title-main">SCN7A</span> Protein-coding gene in the species Homo sapiens

Nax is a protein that in humans is encoded by the SCN7A gene. It is a sodium channel alpha subunit expressed in the heart, the uterus and in glial cells of mice. It has low similarity to all nine other sodium channel alpha subunits (Nav1.1–1.9).

Na<sub>v</sub>1.8 Protein-coding gene in the species Homo sapiens

Nav1.8 is a sodium ion channel subtype that in humans is encoded by the SCN10A gene.

<span class="mw-page-title-main">Gating (electrophysiology)</span>

In electrophysiology, the term gating refers to the opening (activation) or closing of ion channels. This change in conformation is a response to changes in transmembrane voltage.

A depolarizing prepulse (DPP) is an electrical stimulus that causes the potential difference measured across a neuronal membrane to become more positive or less negative, and precedes another electrical stimulus. DPPs may be of either the voltage or current stimulus variety and have been used to inhibit neural activity, selectively excite neurons, and increase the pain threshold associated with electrocutaneous stimulation.

<span class="mw-page-title-main">Ball and chain inactivation</span> Model in neuroscience

In neuroscience, ball and chain inactivation is a model to explain the fast inactivation mechanism of voltage-gated ion channels. The process is also called hinged-lid inactivation or N-type inactivation. A voltage-gated ion channel can be in three states: open, closed, or inactivated. The inactivated state is mainly achieved through fast inactivation, by which a channel transitions rapidly from an open to an inactivated state. The model proposes that the inactivated state, which is stable and non-conducting, is caused by the physical blockage of the pore. The blockage is caused by a "ball" of amino acids connected to the main protein by a string of residues on the cytoplasmic side of the membrane. The ball enters the open channel and binds to the hydrophobic inner vestibule within the channel. This blockage causes inactivation of the channel by stopping the flow of ions. This phenomenon has mainly been studied in potassium channels and sodium channels.

LmαTX5 is an α-scorpion toxin which inhibits the fast inactivation of voltage-gated sodium channels. It has been identified through transcriptome analysis of the venom gland of Lychas mucronatus, also known as the Chinese swimming scorpion – a scorpion species which is widely distributed in Southeast Asia.

References

  1. Yu FH, Catterall WA (2003). "Overview of the voltage-gated sodium channel family". Genome Biology. 4 (3): 207. doi: 10.1186/gb-2003-4-3-207 . PMC   153452 . PMID   12620097.
  2. Nicholls, Martin, Fuchs, Brown, Diamond, Weisblat. (2012) "From Neuron to Brain," 5th ed. pg. 86
  3. Hille, B. (1977) Local Anesthetics: Hydrophilic and Hydrophobic Pathways for the Drug-Receptor Reaction. The Journal of General Physiology, 69, 497-515. http://dx.doi.org/10.1085/jgp.69.4.497
  4. Gamal El-Din, Tamer M., et al. "Fenestrations control resting-state block of a voltage-gated sodium channel." Proceedings of the National Academy of Sciences 115.51 (2018): 13111-13116. https://doi.org/10.1073/pnas.1814928115
  5. Tao, Elaine, and Ben Corry. "Characterizing fenestration size in sodium channel subtypes and their accessibility to inhibitors." Biophysical Journal 121.2 (2022): 193-206. https://doi.org/10.1016/j.bpj.2021.12.025
  6. Yan, Zhen, et al. "Structure of the Nav1. 4-β1 complex from electric eel." Cell 170.3 (2017): 470-482. https://doi.org/10.1016/j.cell.2017.06.039
  7. 1 2 3 Isom LL (February 2001). "Sodium channel beta subunits: anything but auxiliary". The Neuroscientist. 7 (1): 42–54. doi:10.1177/107385840100700108. PMID   11486343. S2CID   86422657.
  8. 1 2 Catterall WA, Goldin AL, Waxman SG (December 2005). "International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels". Pharmacological Reviews. 57 (4): 397–409. doi:10.1124/pr.57.4.4. PMID   16382098. S2CID   7332624.
  9. Noland, Cameron L.; Chua, Han Chow; Kschonsak, Marc; Heusser, Stephanie Andrea; Braun, Nina; Chang, Timothy; Tam, Christine; Tang, Jia; Arthur, Christopher P.; Ciferri, Claudio; Pless, Stephan Alexander; Payandeh, Jian (17 March 2022). "Structure-guided unlocking of NaX reveals a non-selective tetrodotoxin-sensitive cation channel". Nature Communications. 13 (1): 1416. doi:10.1038/s41467-022-28984-4. PMC   8931054 . PMID   35301303.
  10. Lossin C. "SCN1A infobase". Archived from the original on 2011-07-21. Retrieved 2009-10-30. compilation of genetic variations in the SCN1A gene that alter the expression or function of Nav1.1
  11. 1 2 3 4 Bennett DL, Clark AJ, Huang J, Waxman SG, Dib-Hajj SD (April 2019). "The Role of Voltage-Gated Sodium Channels in Pain Signaling". Physiological Reviews. 99 (2): 1079–1151. doi: 10.1152/physrev.00052.2017 . PMID   30672368.
  12. Smith RS, Kenny CJ, Ganesh V, Jang A, Borges-Monroy R, Partlow JN, et al. (September 2018). "V1.3) Regulation of Human Cerebral Cortical Folding and Oral Motor Development". Neuron. 99 (5): 905–913.e7. doi: 10.1016/j.neuron.2018.07.052 . PMC   6226006 . PMID   30146301.
  13. Chockalingam P, Wilde A (September 2012). "The multifaceted cardiac sodium channel and its clinical implications". Heart. 98 (17): 1318–24. doi:10.1136/heartjnl-2012-301784. PMID   22875823. S2CID   44433455.
  14. Beyder A, Mazzone A, Strege PR, Tester DJ, Saito YA, Bernard CE, Enders FT, Ek WE, Schmidt PT, Dlugosz A, Lindberg G, Karling P, Ohlsson B, Gazouli M, Nardone G, Cuomo R, Usai-Satta P, Galeazzi F, Neri M, Portincasa P, Bellini M, Barbara G, Camilleri M, Locke GR, Talley NJ, D'Amato M, Ackerman MJ, Farrugia G (June 2014). "Loss-of-function of the voltage-gated sodium channel NaV1.5 (channelopathies) in patients with irritable bowel syndrome". Gastroenterology. 146 (7): 1659–1668. doi:10.1053/j.gastro.2014.02.054. PMC   4096335 . PMID   24613995.
  15. Butler KM, da Silva C, Shafir Y, Weisfeld-Adams JD, Alexander JJ, Hegde M, Escayg A (January 2017). "De novo and inherited SCN8A epilepsy mutations detected by gene panel analysis". Epilepsy Research. 129: 17–25. doi:10.1016/j.eplepsyres.2016.11.002. PMC   5321682 . PMID   27875746.
  16. Meisler MH, Kearney JA (August 2005). "Sodium channel mutations in epilepsy and other neurological disorders". The Journal of Clinical Investigation. 115 (8): 2010–7. doi:10.1172/JCI25466. PMC   1180547 . PMID   16075041.
  17. Vargas-Alarcon G, Alvarez-Leon E, Fragoso JM, Vargas A, Martinez A, Vallejo M, Martinez-Lavin M (February 2012). "A SCN9A gene-encoded dorsal root ganglia sodium channel polymorphism associated with severe fibromyalgia". BMC Musculoskeletal Disorders. 13 (1): 23. doi: 10.1186/1471-2474-13-23 . PMC   3310736 . PMID   22348792.
  18. Catterall WA (April 2000). "From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels". Neuron. 26 (1): 13–25. doi: 10.1016/S0896-6273(00)81133-2 . PMID   10798388. S2CID   17928749.
  19. Isom LL, De Jongh KS, Patton DE, Reber BF, Offord J, Charbonneau H, Walsh K, Goldin AL, Catterall WA (May 1992). "Primary structure and functional expression of the beta 1 subunit of the rat brain sodium channel". Science. 256 (5058): 839–42. Bibcode:1992Sci...256..839I. doi:10.1126/science.1375395. PMID   1375395.
  20. "Blackboard Server Unavailable" (PDF). blackboard.jhu.edu. Retrieved 2020-07-16.
  21. Malhotra JD, Kazen-Gillespie K, Hortsch M, Isom LL (April 2000). "Sodium channel beta subunits mediate homophilic cell adhesion and recruit ankyrin to points of cell-cell contact". The Journal of Biological Chemistry. 275 (15): 11383–8. doi: 10.1074/jbc.275.15.11383 . PMID   10753953.
  22. Malhotra JD, Koopmann MC, Kazen-Gillespie KA, Fettman N, Hortsch M, Isom LL (July 2002). "Structural requirements for interaction of sodium channel beta 1 subunits with ankyrin". The Journal of Biological Chemistry. 277 (29): 26681–8. doi: 10.1074/jbc.M202354200 . PMID   11997395.
  23. Cantrell AR, Catterall WA (June 2001). "Neuromodulation of Na+ channels: an unexpected form of cellular plasticity". Nature Reviews. Neuroscience. 2 (6): 397–407. doi:10.1038/35077553. PMID   11389473. S2CID   22885909.
  24. Shah BS, Rush AM, Liu S, Tyrrell L, Black JA, Dib-Hajj SD, Waxman SG (August 2004). "Contactin associates with sodium channel Nav1.3 in native tissues and increases channel density at the cell surface". The Journal of Neuroscience. 24 (33): 7387–99. doi:10.1523/JNEUROSCI.0322-04.2004. PMC   6729770 . PMID   15317864.
  25. Wittmack EK, Rush AM, Craner MJ, Goldfarb M, Waxman SG, Dib-Hajj SD (July 2004). "Fibroblast growth factor homologous factor 2B: association with Nav1.6 and selective colocalization at nodes of Ranvier of dorsal root axons". The Journal of Neuroscience. 24 (30): 6765–75. doi:10.1523/JNEUROSCI.1628-04.2004. PMC   6729706 . PMID   15282281.
  26. Rush AM, Wittmack EK, Tyrrell L, Black JA, Dib-Hajj SD, Waxman SG (May 2006). "Differential modulation of sodium channel Na(v)1.6 by two members of the fibroblast growth factor homologous factor 2 subfamily". The European Journal of Neuroscience. 23 (10): 2551–62. doi:10.1111/j.1460-9568.2006.04789.x. PMID   16817858. S2CID   21411801.
  27. Kazarinova-Noyes K, Malhotra JD, McEwen DP, Mattei LN, Berglund EO, Ranscht B, Levinson SR, Schachner M, Shrager P, Isom LL, Xiao ZC (October 2001). "Contactin associates with Na+ channels and increases their functional expression". The Journal of Neuroscience. 21 (19): 7517–25. doi:10.1523/JNEUROSCI.21-19-07517.2001. PMC   6762905 . PMID   11567041.
  28. Srinivasan J, Schachner M, Catterall WA (December 1998). "Interaction of voltage-gated sodium channels with the extracellular matrix molecules tenascin-C and tenascin-R". Proceedings of the National Academy of Sciences of the United States of America. 95 (26): 15753–7. Bibcode:1998PNAS...9515753S. doi: 10.1073/pnas.95.26.15753 . PMC   28116 . PMID   9861042.
  29. Kim DY, Ingano LA, Carey BW, Pettingell WH, Kovacs DM (June 2005). "Presenilin/gamma-secretase-mediated cleavage of the voltage-gated sodium channel beta2-subunit regulates cell adhesion and migration". The Journal of Biological Chemistry. 280 (24): 23251–61. doi: 10.1074/jbc.M412938200 . PMID   15833746.
  30. Ratcliffe CF, Westenbroek RE, Curtis R, Catterall WA (July 2001). "Sodium channel beta1 and beta3 subunits associate with neurofascin through their extracellular immunoglobulin-like domain". The Journal of Cell Biology. 154 (2): 427–34. doi:10.1083/jcb.200102086. PMC   2150779 . PMID   11470829.
  31. 1 2 3 Antzelevitch C, Patocskai B (January 2016). "Brugada Syndrome: Clinical, Genetic, Molecular, Cellular, and Ionic Aspects". Current Problems in Cardiology. 41 (1): 7–57. doi:10.1016/j.cpcardiol.2015.06.002. PMC   4737702 . PMID   26671757.
  32. Moran Y, Barzilai MG, Liebeskind BJ, Zakon HH (February 2015). "Evolution of voltage-gated ion channels at the emergence of Metazoa". The Journal of Experimental Biology. 218 (Pt 4): 515–25. doi: 10.1242/jeb.110270 . PMID   25696815.
  33. Liebeskind BJ, Hillis DM, Zakon HH (May 2011). "Evolution of sodium channels predates the origin of nervous systems in animals". Proceedings of the National Academy of Sciences of the United States of America. 108 (22): 9154–9. Bibcode:2011PNAS..108.9154L. doi: 10.1073/pnas.1106363108 . PMC   3107268 . PMID   21576472.
  34. King N, Westbrook MJ, Young SL, Kuo A, Abedin M, Chapman J, et al. (February 2008). "The genome of the choanoflagellate Monosiga brevicollis and the origin of metazoans". Nature. 451 (7180): 783–8. Bibcode:2008Natur.451..783K. doi:10.1038/nature06617. PMC   2562698 . PMID   18273011.
  35. 1 2 Strong M, Chandy KG, Gutman GA (January 1993). "Molecular evolution of voltage-sensitive ion channel genes: on the origins of electrical excitability". Molecular Biology and Evolution. 10 (1): 221–42. doi: 10.1093/oxfordjournals.molbev.a039986 . PMID   7680747.
  36. Liebeskind BJ, Hillis DM, Zakon HH (November 2013). "Independent acquisition of sodium selectivity in bacterial and animal sodium channels". Current Biology. 23 (21): R948–9. doi: 10.1016/j.cub.2013.09.025 . PMID   24200318.
  37. Kasimova MA, Granata D, Carnevale V (2016). "Voltage-Gated Sodium Channels: Evolutionary History and Distinctive Sequence Features". Current Topics in Membranes. Vol. 78. pp. 261–86. doi:10.1016/bs.ctm.2016.05.002. ISBN   9780128053867. PMID   27586287.
  38. 1 2 Widmark J, Sundström G, Ocampo Daza D, Larhammar D (January 2011). "Differential evolution of voltage-gated sodium channels in tetrapods and teleost fishes". Molecular Biology and Evolution. 28 (1): 859–71. doi: 10.1093/molbev/msq257 . PMID   20924084.
  39. 1 2 Zakon HH, Jost MC, Lu Y (April 2011). "Expansion of voltage-dependent Na+ channel gene family in early tetrapods coincided with the emergence of terrestriality and increased brain complexity". Molecular Biology and Evolution. 28 (4): 1415–24. doi:10.1093/molbev/msq325. PMC   3058772 . PMID   21148285.
  40. 1 2 Zakon H, Jost M, Zwickl D, Lu Y, Hillis D (2009). "Molecular evolution of Na+ channels in teleost fishes". Integrative Zoology. 4 (1): 64–74. doi:10.1111/j.1749-4877.2008.00136.x. PMID   21392277.