Mechanosensitive channels

Last updated

Mechanosensitive channels (MSCs), mechanosensitive ion channels or stretch-gated ion channels are membrane proteins capable of responding to mechanical stress over a wide dynamic range of external mechanical stimuli. [1] [2] [3] [4] They are present in the membranes of organisms from the three domains of life: bacteria, archaea, and eukarya. [5] They are the sensors for a number of systems including the senses of touch, hearing and balance, as well as participating in cardiovascular regulation and osmotic homeostasis (e.g. thirst). The channels vary in selectivity for the permeating ions from nonselective between anions and cations in bacteria, to cation selective allowing passage Ca2+, K+ and Na+ in eukaryotes, and highly selective K+ channels in bacteria and eukaryotes.

Contents

All organisms, and apparently all cell types, sense and respond to mechanical stimuli. [6] MSCs function as mechanotransducers capable of generating both electrical and ion flux signals as a response to external or internal [7] stimuli. [8] Under extreme turgor in bacteria, non selective MSCs such as MSCL and MSCS serve as safety valves to prevent lysis. In specialized cells of the higher organisms, other types of MSCs are probably the basis of the senses of hearing and touch and sense the stress needed for muscular coordination. However, none of these channels have been cloned. MSCs also allow plants to distinguish up from down by sensing the force of gravity. MSCs are not pressure-sensitive, but sensitive to local stress, most likely tension in the surrounding lipid bilayer. [9]

History

Mechanosensitive channels were discovered in 1983 in the skeletal muscle of embryonic chicks [10] by Falguni Guharay and Frederick Sachs. [11] They were also observed (pub. 1986) in Xenopus oocytes, [12] and frequently studied since that time. [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] Since then, MSCs have been found in cells from bacteria to humans: [24] they are now known to be present in all three domains of life (Archaea, Bacteria and Eukarya, incl. plants and fungi). [25] In the decades since the discovery of MS, the understanding of their structure and function has increased greatly, and several have been cloned. Specifically, the cloned eukaryotic mechanosensitive channels include the K+ selective 2P domain channels [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] and the recently cloned cation selective PIEZO family (PIEZO1 and PIEZO2). [40] [41] [42] [43] [44] [45]

Classification

MSCs can be classified based on the type of ion to which they are permeable:

Broadly, most MSCs can be classified as lipid-gated channels.

Functions

For a protein to be considered mechanosensitive, it must respond to a mechanical deformation of the membrane. Mechanical deformations can include changes in the tension, thickness, or curvature of the membrane. Mechanosensitive channels respond to membrane tension by altering their conformation between an open state and a closed state. [48] [49] One type of mechanically sensitive ion channel activates specialized sensory cells, such as cochlear hair cells and some touch sensory neurons, in response to forces applied to proteins. [50] [51]

Stretch-activated ion channels are required for the initial formation of an action potential from a mechanical stimulus, for example by the mechanoreceptors in vibrissae (whiskers) of some animals such as rodents.

Afferent nerve fibers responsible for sensory stimulus detection and feedback are especially sensitive to stimulation. This results from the specialized mechanoreceptor cells that are superimposed upon the afferent nerve fibers. Stretch-activated ion channels are located on these mechanoreceptor cells and serve to lower the action potential threshold, thus making the afferent nerves more sensitive to stimulation. Afferent nerve endings without mechanoreceptor cells are called free nerve endings. They are less sensitive than the encapsulated afferent fibers and generally function in the perception of pain. [52]

Stretch-activated ion channels are responsible for many bodily functions in mammals. In the skin they are responsible for sensing vibration, pressure sensation, stretch, touch, and light touch. [53] [54] They are expressed in sensory modalities including taste, hearing, smell, heat sensation, volume control, and vision. [55] [56] [57] They can also regulate internal functions of our body including, but not limited to, osmotic pressure in cells, blood pressure in veins and arteries, micturition, and heart electrophysiology [58] [59] and contractility. [55] [57] In addition to these functionalities, stretch-activated ion channels have also been found to be involved with balance and proprioceptive sensation. [55]

Channels that have traditionally been known as just "voltage-" or "ligand-gated" have also been found to be mechanically sensitive as well. Channels exhibit mechanical sensitivity as a general property. However, mechanical stress affects various types of channels in different ways. Voltage and ligand gated channels can be modified slightly by mechanical stimulation, which might change their responsiveness or permeability slightly, but they still respond primarily to voltage or ligands, respectively. [60]

Examples

The different families of stretch-activated ion channels are responsible for different functions around the body. The DEG/ENaC family consists of two subgroups: the ENaC subfamily regulates Na+ reabsorption in kidney and lung epithelia; the ASIC subfamily is involved in fear conditioning, memory formation, and pain sensation. [61] The TRP superfamily of channels are found in sensory receptor cells that are involved in heat sensation, taste, smell, touch, and osmotic and volume regulation. [56] MscM, MscS, and MscL channels (mechanosensitive channels of mini, small, and large conductance) regulate osmotic pressure in cells by releasing intracellular fluid when they become too stretched. [55] In the body, a possible role in myoblast development has been described. [62] Furthermore, mechanically gated ion channels are also found in the stereocilia of the inner ear. Sound waves are able to bend the stereocilia and open up ion channels leading to the creation of nerve impulses. [63] These channels also play a role in sensing vibration and pressure via activation of Pacinian corpuscles in the skin. [64]

Transduction mechanisms

There are two different types of stretch-activated channels between which it is important to distinguish: mechanically gated channels, which are directly influenced by mechanical deformations of the membrane, and mechanically sensitive channels, which are opened by second messengers released from the true mechanically gated channel. [53]

Mechanical deformations in the cell membrane can increase the probability of the channels opening. Proteins of the extracellular matrix and cytoskeleton are tethered to extra - and intra-cytoplasmic domains, respectively, of the stretch-activated ion channels. Tension on these mechanosensory proteins causes these proteins to act as a signaling intermediate, resulting in the opening of the ion channel. [53] All known stretch-activated ion channels in prokaryotic cells have been found to be opened by direct deformation of the lipid bilayer membrane. [55] Channels that have been shown to exclusively use this mechanism of gating are the TREK-1 and TRAAK channels. In studies using mammalian hair cells, the mechanism that pulls on proteins tethered from the intra- and extra-cytoplasmic domain of the channel to the cytoskeleton and extracellular matrix, respectively, is the most likely model for ion channel opening. [55]

Mechanical deformation of the cell membrane can be achieved by a number of experimental interventions, including magnetic actuation of nanoparticles. An example of this is the control of calcium influx of axons and boutons within neural networks. [65] Note that this is not an indication of 'magnetic stimulation' of mechanosensitive channels.

Gating mechanism

Although MS vary in many aspects, structures and functions, all the MS studied to date share an important feature: in a process called gating, they all open in a pore-like manner when protein channels are activated by a mechanical stimulus. There are currently two models of the gating process that explain how membrane-activated ion channels open.

Gating Mechanism of MS.Stretch activated model, tension in the lipid bilayer triggers conformational changes which open the channel. Figure adapted from Lumpkin et al. Stretch model,.jpg
Gating Mechanism of MS.Stretch activated model, tension in the lipid bilayer triggers conformational changes which open the channel. Figure adapted from Lumpkin et al.

Lipid bilayer Tension or stretch model: [68] In this model tension in the lipid bilayer triggers conformational changes, thus leading to the opening of the channels. The tension perceived by the protein comes from the lipids. It has been demonstrated that the tension/stretch profile in the lipid bilayer is originated by membrane curvature and bilayer-protein hydrophobic mismatch. [69]

Gating Mechanism of MSC:Spring-like tether model - The tethers are attached to the channel proteins and are connected to the cytoskeleton. The tethers act like spring mechanisms of a shutter. Figure adapted from Lumpkin et al. Spring-like model.jpg
Gating Mechanism of MSC:Spring-like tether model - The tethers are attached to the channel proteins and are connected to the cytoskeleton. The tethers act like spring mechanisms of a shutter. Figure adapted from Lumpkin et al.

Spring-like Tether model: In this model a spring-like tether is attached directly to the MS channel and can be present in either the cytoskeleton or the extracellular matrix linking these elements together. When external stimuli deflect the tether the displacement opens the channel. [67] This particular mechanism has been demonstrated to be the responsible for gating hair cells which are responsible for hearing in vertebrates. [70]

Bacterial MSCs

Bacterial MS channels were first discovered by patch-clamp experiments in E. coli. [71] They have been classified based on their conductance as mini (MscM), small (MscS) and large large (MscL)). These channels function in tandem-mode and are responsible of turgor regulation in bacteria; when activated by changes in the osmotic pressure. MscM is activated first at really low pressures followed by MscS, and finally MscL being the last chance of survival during osmotic shock. Their task was demonstrated when bacteria missing both MscS and MscL were lysed after exposure to osmotic downshocks. [72]

MscS: Small conductance mechanosensitive channel.

The closed structure of MscS MscS close state.jpg
The closed structure of MscS

The main conductance is 1nS in buffer solution. Channel-proteins have been found in gram-positive and gram-negative bacteria, archaea and plants. MscS channel was found after studies in E. coli spheroplasts. [69] The identification of the gene family necessitated for MS of small conductance was as two different channels. YggB encoding MscS and KefA encoding MscK in E. coli further confirm its role osmotic regulation. Mutagenesis studies showed that when both genes YggB and KefA were deleted MscS lost its function, but maintain MscL and MscM, but mutants deficient of YggB and MscL showed that the function of those channel is to open in respond to pressure range right before cell rupture. [73]

The 3D structure of this channel at closed state was elucidated after the crystallography study by Bass et al. [74] which showed that at resolution of 3.9 Å this 31kDa protein is an homoheptamer forming a channel with 80 Å of diameter and 120 Å in length, each subunit contains three transmembrane domains (TM1, TM2, and TM3) with the N-terminal facing the periplasm and the C-terminal embedded in the cytoplasm. The TM3 is highly conserved in MscS family and it is thought to play an important role in MS prokaryotic gating. [75] MscS is a small protein composed of 286 amino acid residues activated by both tension in the lipid bilayer and voltage; in 2002 Vasquez et al. [76] detailed this process and showed that during the change from closed state to open state the TM1 tilt and rotate making TM2 being exposed to the membrane and the TM3 helices expand, tilt, and rotate. During the rearrangement the confined part of the pore was measured as 11 Å, and water molecules were more accessible to the TM3. The two transmembrane domains are in continuous contact with the lipid bilayer and are thought to be the sensor for the tension in the lipid bilayer as well as sensor for voltage because of the three arginine residues present in those domains. [77]

Although MscS is activated by voltage it has been demonstrated that, voltage itself is insufficient to open the channel, thus functioning in a cooperative manner with the channel. The more positive voltage, the higher the probabilities of opening the channel as long as pressure over the threshold is still applied in the system; the performance of this channel at higher voltage has not been completely understood. MscS has a small affinity for negative ions including Cl-, and glutamate. [78]

MscL: Large conductance mechanosensitive channel.

The closed structure of MscL 2oar.jpg
The closed structure of MscL

In bacteria MscL was the first MS channels cloned and sequenced, and is by far one of the most studied channels. The gene encoding MscL protein is trkA and it is located in the inner membrane of the E. coli. The protein is 17 KDa, and consists of 136 amino acids; mostly hydrophobic residues resulting in two hydrophobic segments, however molecular weight of the functional channel is presumed to be 60-70 KDa from gel filtration experiments, suggesting oligomerization. As a common feature no cysteines residues are present in this channel. [79]

In 1998 the homolog MscL from mycobacterium tuberculosis Tb-MscL was elucidated at closed state by X ray crystallography at 3.5 Å resolution. The protein is a homopentamer composed mostly of helical regions trans orientation of the helices with respect to the bilayer, with two domains: the cytoplasmic and the transmembrane. The channel is 85 Å in length, 35 Å and 50 Å for the cytoplasmic transmembrane domain respectively and 50 Å in diameter. The helices cross the membrane twice with both the C-terminal and the N-terminal, thus having two transmembrane domains TM1 and TM2 being TM1 the most conserved region among MscL proteins especially at the N-terminal region. [80] It is located in the cytoplasm and forms a α-hydrophobic helix called S1; the region between the transmembrane domains form a loop that is divided into two regions: S2 a glycine-proline rich region and S3 a short helical section. [81] The secondary structure of the protein is resistant to thermal denaturation still in the presence of SDS. [82]

During the activation of the prokaryotic MscL by tension in the lipid bilayer an intermediate state was determined. The S1 segments form a bundle when the structure is in the closed state, and the crosslinking of S1 segments prevents the opening of the channel. When tension is applied to the membrane the transmembrane barrel-like structure expand and stretch apart the region S1-TM1 allowing the channel to open. [83] The size of the pore at open state is approximately 25Å. The transition from closed to intermediate state is accompanied by small movements of the TM1; further transitions to the open stated are characterized by big rearrangements in both the TM1 and TM2. [84]

Role of lipid bilayer in MS

The lipid bilayer is an important structure in all living cells; it has many functions such as separation of compartments, and signaling among others. In the case of the prokaryotic protein channels MscS and MscL both are gated by tension in the lipid bilayer, thus suggesting an important role in such a complex structures.

The tension in the membrane bilayer has been extensively studied, simple intrinsic properties of the lipids can account for the contributions in the free energy of the open, intermediate, and close state of the MS channels. The bilayer possess different features that allows it to transduce tension and to prevent exhaustive deformations, the first one is "in plane fluidity of the lipid bilayer" meaning that any in plane tension in the lipid bilayer is felt homogenously in the absence of cytoskeleton interactions. The lipid molecules have specific spaces in between them which prevent changes in the lipid bilayer. [85]

The contribution of membrane deformation in the gating of MS channels can be divided in two types: the deformation of the plane of the bilayer, and the deformation of the thickness of the bilayer. Also during any process involving changes in the structure, the free energy of the process itself is also an important factor. During gating the major processes that account for this event are: hydrophobic mismatch, and membrane curvature. It has been calculated that the free energy of the tension in the lipid bilayer is similar to the energy needed for gating the channels. [86]

A different study showed that the length of the hydrophobic tail affects its functioning as well as supporting the different states, phosphatidylcholine (PC) 18 stabilizes better the open state of the MscL channel, PC 14 stabilizes the intermediate state, and a mixture of PC 18 and lysophosphatidylcholine (LPC) stabilizes the closed state, [84] suggesting that the bilayer thickness (for carbon tail lengths of 16, 18 and 20) affects channel function. In conclusion the energy from the environment of the membrane plays an important role in the total energy of channel gating.

Eukaryotes

In eukaryotes, two of the best known mechanosensitive ion channels are the potassium channels TREK-1 and TRAAK, both of which are found in mammalian neurons.

Recently, a new mechanosensitive ion channel family was cloned, with two mammalian members, PIEZO1 and PIEZO2. [87] Both these channels are expressed in the lungs and bladder, organs with important mechanosensory functions. Piezo1 is also expressed in the skin, and in red blood cells, and its gain of function mutations cause hereditary xerocytosis. [88] Piezo2 is expressed in sensory neurons of the dorsal root and trigeminal ganglia indicating that it may play a role in touch sensation. Mutations in piezo2 are associated with a human disease named Distal Arthrogryposis. [89]

Physiological role of MS

MS channels are ubiquitously expressed in the membrane of prokaryotes suggesting their significance. In Bacteria and Archaea the function of these channels is conserved and it has been demonstrated that they play a role in turgor regulation. In Eukarya MS channels are involved in all five senses. The main family is TRP, and one good example is hair cells involved in the hearing process. When a wave of sound deflects the stereocilia, the channel opens. This is an instance of the Spring-like Tether gating mechanism. Recent studies have revealed a new role of mechanosensitive pathways in which naive mesenchymal stem cells are committed to a particular lineage based on the elasticity of its surrounding matrix. [90]

Some MS channels that have been cloned and characterized. Data adapted from Martinac, 2001 [91]
ChannelSourceGating mechanismPhysiological role
MscLBacteriaLipid bilayerTurgor regulation and cell growth
MscSBacteriaLipid bilayerTurgor regulation and cell growth
MscMJArchaeaLipid bilayerTurgor regulation
MEC4C. elegansTetherTouch
TRPYFungiBilayerTurgor regulation
TRECK-1MammalianBilayerResting membrane potential

MS have also been suggested as a potential target for antibiotics, the reasoning behind this idea is that both McsS and MscL are highly conserved among prokaryotes, but their homologs have not been found in animals [92] making them an exceptional potential for further studies.

In mammalian neurons, opening of the ion channels depolarizes the afferent neuron producing an action potential with sufficient depolarization. [52] Channels open in response to two different mechanisms: the prokaryotic model and the mammalian hair cell model. [55] [56] Stretch-activated ion channels have been shown to detect vibration, pressure, stretch, touch, sounds, tastes, smell, heat, volume, and vision. [53] [54] [57] Stretch-activated ion channels have been categorized into three distinct "superfamilies": the ENaC/DEG family, the TRP family, and the K1 selective family. These channels are involved with bodily functions such as blood pressure regulation. [60] They are shown to be associated with many cardiovascular diseases. [56] Stretch-activated channels were first observed in chick skeletal muscles by Falguni Guharay and Frederick Sachs in 1983 and the results were published in 1984. [93] Since then stretch-activated channels have been found in cells from bacteria to humans as well as plants.

The opening of these channels is central to a neuron's response to pressure, often osmotic pressure and blood pressure, to regulate ionic flow in internal environments. [55]

Techniques used to study MS

This is a short list of the most frequently techniques used to study the properties, function, mechanism and other features of these channels:

Finite Element Model of MscL, a bacterial channel. This figure is similar to that in Tang et al. Finite Element Model.jpg
Finite Element Model of MscL, a bacterial channel. This figure is similar to that in Tang et al.

Through experiments performed on the cytoskeleton and extra-cytoplasmic matrix of stretch-activated ion channels, these structures have been shown to play significant roles in mechanotransduction. [53] In one such experiment on adult heart cells, whole cell recordings were taken on cells being squeezed with two pipettes at 1 Hz/1 um. This squeezing produced no current until five minutes in when a large depolarization was observed. Hereafter, the cell became extremely responsive to every compression and gradually decreased sensitivity over the next few minutes. [60] Researchers hypothesized that, initially, the cytoskeleton was buffering the mechanical deformation of the squeezing from the channel. The depolarization at five minutes was the cytoskeleton snapping which subsequently caused the channel to sense the mechanical deformations and thereby respond to the stimuli. Researchers believe that over the few minutes where the channel repaired itself the cytoskeleton must be repairing itself and newly adapting to the squeezing stimuli. [60]

Structure

ENaC/DEG superfamily

ASIC

There are six known ASIC subunits, ASIC1a, ASIC1b, ASIC2a, ASIC2b, ASIC3, and ASIC4, which have two transmembrane domains, extracellular and intracellular loops, and C and N termini. These ASIC subunits likely form tetramers with varying kinetics, pH sensitivity, tissue distribution, and pharmacological properties. [53]

TRP superfamily

There are seven subfamilies within the TRP superfamily: TRPC (canonical), TRPV (vanilloid), TRPM (melastatin), TRPP (polycystin), TRPML (mucolipin), TRPA (ankyrin), and TRPN (NOMPC-like). [53] TRP proteins typically consist of six transmembrane domains, S1, S2, S3, S4, S5, and S6, with a pore between S5 and S6. These contain intracellular N and C termini, which form tetramers [61] and vary in length and domain. [53] Within the channel there are ankyrins, which are structural proteins that mediate protein-protein interactions, and are thought to contribute to the tether model of stretch-activated channel opening. NOMPC, identified in D. melanogaster mechanotransduction and a member of the TRPN subfamily, contains a relatively high number of ankyrins. [55]

K1-selective superfamily

K2P channels consist of six subfamilies and contain four transmembrane domains, which form two pores each between domains 1–2 and 3–4. K2P channels also contain a short N terminal domain and a C terminal which varies in length. There is also a large extracellular linker region between domain 1 and the first pore formed between domains 1–2. [53]

Examples

TRP channels are typically non-selective, although a few are selective for calcium or hydrated magnesium ions, and are composed of integral membrane proteins. Although many TRP channels are activated by voltage change, ligand binding, or temperature change, [53] some TRP channels have been hypothesized to be involved in mechanotransduction. [56] Some examples are TRPV4, which mediates mechanical load in a variety of tissues, including the liver, heart, lung, trachea, testis, spleen, salivary glands, cochlea, and vascular endothelial cells, [56] as well as TRPC1 and TRPC6, which are involved in muscle mechanosensation. TRPC1 is expressed in the myocytes of the heart, arteries, and skeletal muscle. TRPC1 is widely considered to be a non-selective "store-operated ion channel" (SOC) involved in the calcium influx following calcium depletion of the endoplasmic reticulum of the cell. [95] TRPC6 is a calcium-permeable non-selective cation channel expressed in the cardiovascular system. TRPC6 is potentially a sensor of mechanically and osmotically induced membrane stretch, and is possibly directly gated by membrane tension. [95] Other examples include TREK-1 and TRAAK which are found in mammalian neurons and are classified as potassium channels in the tandem pore domain class [96] [97] and "MID-1" (also known as "MCLC" or CLCC1.) [98] [99]

The six K2P channel subfamilies are regulated by various physical, cellular, and pharmacological stimulants, including membrane stretch, heat, pH change, calcium flux, and protein kinases. [53]

Clinical relevance

Stretch-activated ion channels perform important functions in many different areas of our body. Pressure-dependent myogenic constriction resistance arteries require these channels for regulation in the smooth muscle of the arteries. [54] They have been found to be used for volume sensing in animals and blood pressure regulation. [60] Bacteria have been shown to relieve hydrostatic pressure through MscL and MscS channels. [60]

Pathologies associated with stretch-activated ion channels

Stretch-activated ion channels have been correlated with major pathologies. Some of these pathologies include cardiac arrhythmia (such as atrial fibrillation), [60] cardiac hypertrophy, Duchenne muscular dystrophy, [54] and other cardiovascular diseases. [56]

Blocking stretch-activated ion channels

Gadolinium (Gd3+) and other lanthanides have been shown to block stretch-activated ion channel function. The peptide toxin isolated from the Chilean rose tarantula (Grammostola rosea, synonym G. spatulata), mechanotoxin 4 (GsMTx4) has been shown to inhibit these channels from the extracellular side, but it does not inhibit all stretch-activated ion channels and particularly has no effect on 2p channels. [60]

List of diseases associated with mechanosensitive channels

Abnormalities in the function of MS channels can cause: [25]

See also

Related Research Articles

<span class="mw-page-title-main">Ion channel</span> Pore-forming membrane protein

Ion channels are pore-forming membrane proteins that allow ions to pass through the channel pore. Their functions include establishing a resting membrane potential, shaping action potentials and other electrical signals by gating the flow of ions across the cell membrane, controlling the flow of ions across secretory and epithelial cells, and regulating cell volume. Ion channels are present in the membranes of all cells. Ion channels are one of the two classes of ionophoric proteins, the other being ion transporters.

<span class="mw-page-title-main">Lipid bilayer</span> Membrane of two layers of lipid molecules

The lipid bilayer is a thin polar membrane made of two layers of lipid molecules. These membranes are flat sheets that form a continuous barrier around all cells. The cell membranes of almost all organisms and many viruses are made of a lipid bilayer, as are the nuclear membrane surrounding the cell nucleus, and membranes of the membrane-bound organelles in the cell. The lipid bilayer is the barrier that keeps ions, proteins and other molecules where they are needed and prevents them from diffusing into areas where they should not be. Lipid bilayers are ideally suited to this role, even though they are only a few nanometers in width, because they are impermeable to most water-soluble (hydrophilic) molecules. Bilayers are particularly impermeable to ions, which allows cells to regulate salt concentrations and pH by transporting ions across their membranes using proteins called ion pumps.

<span class="mw-page-title-main">Transmembrane protein</span> Protein spanning across a biological membrane

A transmembrane protein is a type of integral membrane protein that spans the entirety of the cell membrane. Many transmembrane proteins function as gateways to permit the transport of specific substances across the membrane. They frequently undergo significant conformational changes to move a substance through the membrane. They are usually highly hydrophobic and aggregate and precipitate in water. They require detergents or nonpolar solvents for extraction, although some of them (beta-barrels) can be also extracted using denaturing agents.

<span class="mw-page-title-main">Stimulus (physiology)</span> Detectable change in the internal or external surroundings

In physiology, a stimulus is a detectable change in the physical or chemical structure of an organism's internal or external environment. The ability of an organism or organ to detect external stimuli, so that an appropriate reaction can be made, is called sensitivity (excitability). Sensory receptors can receive information from outside the body, as in touch receptors found in the skin or light receptors in the eye, as well as from inside the body, as in chemoreceptors and mechanoreceptors. When a stimulus is detected by a sensory receptor, it can elicit a reflex via stimulus transduction. An internal stimulus is often the first component of a homeostatic control system. External stimuli are capable of producing systemic responses throughout the body, as in the fight-or-flight response. In order for a stimulus to be detected with high probability, its level of strength must exceed the absolute threshold; if a signal does reach threshold, the information is transmitted to the central nervous system (CNS), where it is integrated and a decision on how to react is made. Although stimuli commonly cause the body to respond, it is the CNS that finally determines whether a signal causes a reaction or not.

<span class="mw-page-title-main">Theories of general anaesthetic action</span> How drugs induce reversible suppression of consciousness

A general anaesthetic is a drug that brings about a reversible loss of consciousness. These drugs are generally administered by an anaesthetist/anesthesiologist to induce or maintain general anaesthesia to facilitate surgery.

<span class="mw-page-title-main">Mechanotransduction</span> Conversion of mechanical stimulus of a cell into electrochemical activity

In cellular biology, mechanotransduction is any of various mechanisms by which cells convert mechanical stimulus into electrochemical activity. This form of sensory transduction is responsible for a number of senses and physiological processes in the body, including proprioception, touch, balance, and hearing. The basic mechanism of mechanotransduction involves converting mechanical signals into electrical or chemical signals.

<span class="mw-page-title-main">Large-conductance mechanosensitive channel</span> Group of transport proteins

Large conductance mechanosensitive ion channels (MscLs) (TC# 1.A.22) are a family of pore-forming membrane proteins that are responsible for translating stresses at the cell membrane into an electrophysiological response. MscL has a relatively large conductance, 3 nS, making it permeable to ions, water, and small proteins when opened. MscL acts as stretch-activated osmotic release valve in response to osmotic shock.

Light-gated ion channels are a family of ion channels regulated by electromagnetic radiation. Other gating mechanisms for ion channels include voltage-gated ion channels, ligand-gated ion channels, mechanosensitive ion channels, and temperature-gated ion channels. Most light-gated ion channels have been synthesized in the laboratory for study, although two naturally occurring examples, channelrhodopsin and anion-conducting channelrhodopsin, are currently known. Photoreceptor proteins, which act in a similar manner to light-gated ion channels, are generally classified instead as G protein-coupled receptors.

<span class="mw-page-title-main">KCNK2</span> Protein-coding gene in the species Homo sapiens

Potassium channel subfamily K member 2, also known as TREK-1, is a protein that in humans is encoded by the KCNK2 gene.

<span class="mw-page-title-main">KCNK4</span> Protein-coding gene in the species Homo sapiens

Potassium channel subfamily K member 4 is a protein that in humans is encoded by the KCNK4 gene. KCNK4 protein channels are also called TRAAK channels.

Mechanosensation is the transduction of mechanical stimuli into neural signals. Mechanosensation provides the basis for the senses of light touch, hearing, proprioception, and pain. Mechanoreceptors found in the skin, called cutaneous mechanoreceptors, are responsible for the sense of touch. Tiny cells in the inner ear, called hair cells, are responsible for hearing and balance. States of neuropathic pain, such as hyperalgesia and allodynia, are also directly related to mechanosensation. A wide array of elements are involved in the process of mechanosensation, many of which are still not fully understood.

Lipid bilayer mechanics is the study of the physical material properties of lipid bilayers, classifying bilayer behavior with stress and strain rather than biochemical interactions. Local point deformations such as membrane protein interactions are typically modelled with the complex theory of biological liquid crystals but the mechanical properties of a homogeneous bilayer are often characterized in terms of only three mechanical elastic moduli: the area expansion modulus Ka, a bending modulus Kb and an edge energy . For fluid bilayers the shear modulus is by definition zero, as the free rearrangement of molecules within plane means that the structure will not support shear stresses. These mechanical properties affect several membrane-mediated biological processes. In particular, the values of Ka and Kb affect the ability of proteins and small molecules to insert into the bilayer. Bilayer mechanical properties have also been shown to alter the function of mechanically activated ion channels.

A model lipid bilayer is any bilayer assembled in vitro, as opposed to the bilayer of natural cell membranes or covering various sub-cellular structures like the nucleus. They are used to study the fundamental properties of biological membranes in a simplified and well-controlled environment, and increasingly in bottom-up synthetic biology for the construction of artificial cells. A model bilayer can be made with either synthetic or natural lipids. The simplest model systems contain only a single pure synthetic lipid. More physiologically relevant model bilayers can be made with mixtures of several synthetic or natural lipids.

Mechanobiology is an emerging field of science at the interface of biology, engineering, chemistry and physics. It focuses on how physical forces and changes in the mechanical properties of cells and tissues contribute to development, cell differentiation, physiology, and disease. Mechanical forces are experienced and may be interpreted to give biological responses in cells. The movement of joints, compressive loads on the cartilage and bone during exercise, and shear pressure on the blood vessel during blood circulation are all examples of mechanical forces in human tissues. A major challenge in the field is understanding mechanotransduction—the molecular mechanisms by which cells sense and respond to mechanical signals. While medicine has typically looked for the genetic and biochemical basis of disease, advances in mechanobiology suggest that changes in cell mechanics, extracellular matrix structure, or mechanotransduction may contribute to the development of many diseases, including atherosclerosis, fibrosis, asthma, osteoporosis, heart failure, and cancer. There is also a strong mechanical basis for many generalized medical disabilities, such as lower back pain, foot and postural injury, deformity, and irritable bowel syndrome.

<span class="mw-page-title-main">Cell membrane</span> Biological membrane that separates the interior of a cell from its outside environment

The cell membrane is a biological membrane that separates and protects the interior of a cell from the outside environment. The cell membrane consists of a lipid bilayer, made up of two layers of phospholipids with cholesterols interspersed between them, maintaining appropriate membrane fluidity at various temperatures. The membrane also contains membrane proteins, including integral proteins that span the membrane and serve as membrane transporters, and peripheral proteins that loosely attach to the outer (peripheral) side of the cell membrane, acting as enzymes to facilitate interaction with the cell's environment. Glycolipids embedded in the outer lipid layer serve a similar purpose. The cell membrane controls the movement of substances in and out of a cell, being selectively permeable to ions and organic molecules. In addition, cell membranes are involved in a variety of cellular processes such as cell adhesion, ion conductivity, and cell signalling and serve as the attachment surface for several extracellular structures, including the cell wall and the carbohydrate layer called the glycocalyx, as well as the intracellular network of protein fibers called the cytoskeleton. In the field of synthetic biology, cell membranes can be artificially reassembled.

TRPN is a member of the transient receptor potential channel family of ion channels, which is a diverse group of proteins thought to be involved in mechanoreception. The TRPN gene was given the name no mechanoreceptor potential C (nompC) when it was first discovered in fruit flies, hence the N in TRPN. Since its discovery in fruit flies, TRPN homologs have been discovered and characterized in worms, frogs, and zebrafish.

Small conductance mechanosensitive ion channels (MscS) provide protection against hypo-osmotic shock in bacteria, responding both to stretching of the cell membrane and to membrane depolarization. In eukaryotes, they fulfill a multitude of important functions in addition to osmoregulation. They are present in the membranes of organisms from the three domains of life: bacteria, archaea, fungi and plants.

<span class="mw-page-title-main">PIEZO1</span> Protein-coding gene in the species Homo sapiens

PIEZO1 is a mechanosensitive ion channel protein that in humans is encoded by the gene PIEZO1. PIEZO1 and its close homolog PIEZO2 were cloned in 2010, using an siRNA-based screen for mechanosensitive ion channels.

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

Lipid-gated ion channels are a class of ion channels whose conductance of ions through the membrane depends directly on lipids. Classically the lipids are membrane resident anionic signaling lipids that bind to the transmembrane domain on the inner leaflet of the plasma membrane with properties of a classic ligand. Other classes of lipid-gated channels include the mechanosensitive ion channels that respond to lipid tension, thickness, and hydrophobic mismatch. A lipid ligand differs from a lipid cofactor in that a ligand derives its function by dissociating from the channel while a cofactor typically derives its function by remaining bound.

<span class="mw-page-title-main">GsMTx-4</span> Grammostola mechanotoxin 4

Grammostola mechanotoxin #4, also known as M-theraphotoxin-Gr1a (M-TRTX-Gr1a), is a neurotoxin isolated from the venom of the spider Chilean rose tarantula Grammostola spatulate. This amphiphilic peptide, which consists of 35 amino acids, belongs to the inhibitory cysteine knot (ICK) peptide family. It reduces mechanical sensation by inhibiting mechanosensitive channels (MSCs).

References

  1. Sukharev, S.; Sachs, F. (2012). "Molecular Force Transduction by Ion Channels: diversity and unifying principles". J. Cell Sci. 125 (13): 1–9. doi:10.1242/jcs.092353. PMC   3434843 . PMID   22797911.
  2. Gottlieb, P.; Sachs, F (2012). "The sensation of stretch". Nature. 483 (7388): 163–164. Bibcode:2012Natur.483..163G. doi:10.1038/483163a. PMC   4090763 . PMID   22398551.
  3. Sachs, F. (2010). "Stretch activated Ion Channels; What are They". Physiology. 25 (1): 50–56. doi:10.1152/physiol.00042.2009. PMC   2924431 . PMID   20134028.
  4. Bowman, Charles L.; Gottlieb, P. A.; Suchyna, T. M.; Murphy, Y. K.; Sachs, F. (2007). "Mechanosensitive ion channels and the peptide inhibitor GsMTx-4: History, properties, mechanisms and pharmacology". Toxicon. 49 (2): 249–270. doi:10.1016/j.toxicon.2006.09.030. PMC   1852511 . PMID   17157345.
  5. Pivetti CD, Yen MR, Miller S, Busch W, Tseng YH, Booth IR, Saier MH (March 2003). "Two families of mechanosensitive channel proteins". Microbiol. Mol. Biol. Rev. 67 (1): 66–85, table of contents. doi:10.1128/MMBR.67.1.66-85.2003. PMC   150521 . PMID   12626684.
  6. Kung, C. (2005). "A possible unifying principle for mechanosensation". Nature. 436 (7051): 647–54. Bibcode:2005Natur.436..647K. doi:10.1038/nature03896. PMID   16079835. S2CID   4374012.
  7. Suchyna, T.; Sachs, F. (2007). "Mechanical and electrical properties of membranes from dystrophic and normal mouse muscle". J. Physiol. 581 (Pt 1): 369–387. doi:10.1113/jphysiol.2006.125021. PMC   2075208 . PMID   17255168.
  8. Hackney, CM; Furness, DN (1995). "Mechanotransduction in vertebrate hair cells: structure and function of the stereociliary bundle". Am J Physiol. 268 (1 Pt 1): C1–138. doi:10.1152/ajpcell.1995.268.1.C1. PMID   7840137.
  9. Markin, V. S.; Sachs, F. (2004). "Thermodynamics of mechanosensitivity". Physical Biology. 1 (2): 110–124. Bibcode:2004PhBio...1..110M. doi:10.1088/1478-3967/1/2/007. PMID   16204828. S2CID   24625029.
  10. Guharay, F.; Sachs, F. (July 1984). "Stretch-activated single ion channel currents in tissue-cultured embryonic chick skeletal muscle". J. Physiol. 352: 685–701. doi:10.1113/jphysiol.1984.sp015317. PMC   1193237 . PMID   6086918.
  11. Guharay, F.; Sachs, F. (1985). "Mechanotransducer ion channels in chick skeletal muscle: the effects of extracellular pH". Journal of Physiology. 353: 119–134. doi:10.1113/jphysiol.1985.sp015699. PMC   1192918 . PMID   2410605.
  12. Methfessel, C.; et al. (1986). "Patch clamp measurements on Xenopus laevis oocytes: currents through endogenous channels and implanted acetylcholine receptor and sodium channels". Pflügers Archiv: European Journal of Physiology. 407 (6): 577–588. doi:10.1007/BF00582635. PMID   2432468. S2CID   25200620.
  13. Zhang, Y.; Gao, F.; Popov, V. L.; Wen, J. W.; Hamill, O. P. (2000). "Mechanically gated channel activity in cytoskeleton-deficient plasma membrane blebs and vesicles from Xenopus oocytes". Journal of Physiology. Pt 1. 523 (Pt 1): 117–130. doi:10.1111/j.1469-7793.2000.t01-1-00117.x. PMC   2269789 . PMID   10673548.
  14. Zhang, Y.; Hamill, O. P. (2000). "Calcium-, voltage- and osmotic stress-sensitive currents in Xenopus oocytes and their relationship to single mechanically gated channels". Journal of Physiology. 523 (Pt 1): 83–99. doi:10.1111/j.1469-7793.2000.t01-2-00083.x. PMC   2269778 . PMID   10673546.
  15. Zhang, Y.; Hamill, O. P. (2000). "On the discrepancy between whole-cell and membrane patch mechanosensitivity in Xenopus oocytes". Journal of Physiology. 523 (Pt 1): 101–115. doi:10.1111/j.1469-7793.2000.00101.x. PMC   2269787 . PMID   10673547.
  16. Hamill OP, McBride DW (1997). "Mechanogated channels in Xenopus oocytes: different gating modes enable a channel to switch from a phasic to a tonic mechanotransducer". Biological Bulletin. 192 (1): 121–122. doi:10.2307/1542583. JSTOR   1542583. PMID   9057280.
  17. Hamill, O. P.; McBride, D. W. J. (1996). "Membrane voltage and tension interactions in the gating of the mechano-gated cation channel in xenopus oocytes". Biophysical Journal. 70 (2): A339–A359. Bibcode:1996BpJ....70..339.. doi:10.1016/S0006-3495(96)79669-8. PMC   1225030 .
  18. Wilkinson, N. C.; McBride, D. W.; Hamill, O. P. (1996). "Testing the putative role of a mechano-gated channel in testing Xenopus oocyte maturation, fertilization and tadpole development". Biophysical Journal. 70 (1): 349–357. Bibcode:1996BpJ....70..349Z. doi:10.1016/S0006-3495(96)79576-0. PMC   1224933 . PMID   8770211.
  19. Lane, J. W.; McBride, D. W. Jr; Hamill, O. P. (1993). "Ionic effects on amiloride block of the mechanosensitive channel in Xenopus oocytes". British Journal of Pharmacology. 108 (1): 116–119. doi:10.1111/j.1476-5381.1993.tb13449.x. PMC   1907719 . PMID   7679024.
  20. Hamill, O. P.; McBride, D. W. Jr (1992). "Rapid adaptation of single mechanosensitive channels in Xenopus oocytes". Proceedings of the National Academy of Sciences of the United States of America. 89 (16): 7462–7466. Bibcode:1992PNAS...89.7462H. doi: 10.1073/pnas.89.16.7462 . PMC   49730 . PMID   1380158.
  21. Lane, J. W.; McBride, D. W. Jr; Hamill, O. P. (1992). "Structure-activity relations of amiloride and its analogues in blocking the mechanosensitive channel in Xenopus oocytes". British Journal of Pharmacology. 106 (2): 283–286. doi:10.1111/j.1476-5381.1992.tb14329.x. PMC   1907505 . PMID   1382778.
  22. McBride, D. W. Jr; Hamill, O. P. (1992). "Pressure-clamp: a method for rapid step perturbation of mechanosensitive channels. Pflügers Archiv". European Journal of Physiology. 421 (6): 606–612. doi:10.1007/BF00375058. PMID   1279516. S2CID   27707723.
  23. Lane, J. W.; McBride, D.; Hamill, O. P. (1991). "Amiloride block of the mechanosensitive cation channel in Xenopus oocytes". Journal of Physiology. 441: 347–366. doi:10.1113/jphysiol.1991.sp018755. PMC   1180202 . PMID   1816379.
  24. Sachs, F; Morris, C. E (1998). "Mechanosensitive ion channels in nonspecialized cells". Reviews of Physiology, Biochemistry and Pharmacology. 132: 1–77. doi:10.1007/BFb0004985. ISBN   978-3-540-63492-8. PMID   9558913.
  25. 1 2 "Archived copy" (PDF). Archived from the original (PDF) on 2012-03-17. Retrieved 2012-08-07.{{cite web}}: CS1 maint: archived copy as title (link)
  26. Peyronnet, R. et al. Mechanoprotection by Polycystins against Apoptosis Is Mediated through the Opening of Stretch-Activated K2P Channels. Cell Reports 1 (in press), 241-250 (2012)
  27. Chemin, J.; Patel, AJ; Duprat, F; Sachs, F; Lazdunski, M; Honore, E (2007). "Up- and down-regulation of the mechano-gated K-2P channel TREK-1 by PIP2 and other membrane phospholipids". Pflügers Archiv: European Journal of Physiology. 455 (1): 97–103. doi:10.1007/s00424-007-0250-2. PMID   17384962. S2CID   37929097.
  28. Honore, E. (2007). "The neuronal background K2P channels: focus on TREK1". Nature Reviews Neuroscience. 8 (4): 251–261. doi:10.1038/nrn2117. PMID   17375039. S2CID   21421846.
  29. Chemin, J. et al. in Mechanosensitive Ion Channels, Pt B Vol. 59 Current Topics in Membranes (ed O.P. Hamill) Ch. 7, 155-170 (Academic Press, 2007).>
  30. Honore, E.; Patel, A. J.; Chemin, J.; Suchyna, T.; Sachs, F. (2006). "Desensitization of mechano-gated K-2P channels". Proceedings of the National Academy of Sciences of the United States of America. 103 (18): 6859–6864. Bibcode:2006PNAS..103.6859H. doi: 10.1073/pnas.0600463103 . PMC   1458984 . PMID   16636285.
  31. Chemin, J.; Patel, A; Duprat, F; Zanzouri, M; Lazdunski, M; Honoré, E (2005). "Lysophosphatidic acid-operated K+ channels". Journal of Biological Chemistry. 280 (6): 4415–4421. doi: 10.1074/jbc.M408246200 . PMC   3764821 . PMID   15572365.
  32. Lauritzen, I.; Chemin, J; Honoré, E; Jodar, M; Guy, N; Lazdunski, M; Jane Patel, A (2005). "Cross-talk between the mechano-gated K-2P channel TREK-1 and the actin cytoskeleton". EMBO Reports. 6 (7): 642–648. doi:10.1038/sj.embor.7400449. PMC   1369110 . PMID   15976821.
  33. Honore, E., Patel, A. A., Kohl, P., Franz, M. R. & Sachs, F. in Cardiac Mechano-Electric Feedback and Arrhythmias: From Pipette to Patient (Elsevier 2004)
  34. Maingret F, Honoré E, Lazdunski M, Patel AJ (March 2002). "Molecular basis of the voltage-dependent gating of TREK-1, a mechano-sensitive K(+) channel". Biochem. Biophys. Res. Commun. 292 (2): 339–46. doi:10.1006/bbrc.2002.6674. PMID   11906167.
  35. Patel, A. J.; Lazdunski, M.; Honore, E. (2001). "Lipid and mechano-gated 2P domain K(+) channels". Current Opinion in Cell Biology. 13 (4): 422–428. doi:10.1016/S0955-0674(00)00231-3. PMID   11454447.
  36. Patel, A. J.; Honore, E. (2001). "Properties and modulation of mammalian 2P domain K+ channels". Trends Neurosci. 24 (6): 339–346. doi:10.1016/S0166-2236(00)01810-5. PMID   11356506. S2CID   36875003.
  37. Maingret, F.; Patel, A. J.; Lesage, F.; Lazdunski, M.; Honore, E. (2000). "Lysophospholipids open the two-pore domain mechano-gated K(+) channels TREK-1 and TRAAK". Journal of Biological Chemistry. 275 (14): 10128–10133. doi: 10.1074/jbc.275.14.10128 . PMID   10744694.
  38. Patel, A. J.; Honoré, E; Lesage, F; Fink, M; Romey, G; Lazdunski, M (1999). "Inhalational anesthetics activate two-pore-domain background K+ channels". Nat. Neurosci. 2 (5): 422–426. doi:10.1038/8084. PMID   10321245. S2CID   23092576.
  39. Patel, A. J.; Honoré, E; Maingret, F; Lesage, F; Fink, M; Duprat, F; Lazdunski, M (1998). "A mammalian two pore domain mechano-gated S-like K+ channel". The EMBO Journal. 17 (15): 4283–4290. doi:10.1093/emboj/17.15.4283. PMC   1170762 . PMID   9687497.
  40. Coste, Bertrand; Xiao, Bailong; Santos, Jose S.; Syeda, Ruhma; Grandl, Jörg; Spencer, Kathryn S.; Kim, Sung Eun; Schmidt, Manuela; et al. (2012). "Piezo proteins are pore-forming subunits of mechanically activated channels". Nature. 483 (7388): 176–81. Bibcode:2012Natur.483..176C. doi:10.1038/nature10812. PMC   3297710 . PMID   22343900.
  41. Kim, Sung Eun; Coste, Bertrand; Chadha, Abhishek; Cook, Boaz; Patapoutian, Ardem (2012). "The role of Drosophila Piezo in mechanical nociception". Nature. 483 (7388): 209–12. Bibcode:2012Natur.483..209K. doi:10.1038/nature10801. PMC   3297676 . PMID   22343891.
  42. Coste, B.; Mathur, J.; Schmidt, M.; Earley, T. J.; Ranade, S.; Petrus, M. J.; Dubin, A. E.; Patapoutian, A. (2010). "Are Essential Components of Distinct Mechanically Activated Cation Channels". Science. 330 (6000): 55–60. Bibcode:2010Sci...330...55C. doi:10.1126/science.1193270. PMC   3062430 . PMID   20813920.
  43. Gottlieb, P.; Sachs, F. Piezo (2012). "Properties of a cation selective mechanical channel". Channels. 6 (4): 1–6. doi:10.4161/chan.21050. PMC   3508900 . PMID   22790400.
  44. Gottlieb, P. A.; Sachs, F. (2012). "CELL BIOLOGY The sensation of stretch". Nature. 483 (7388): 163–164. Bibcode:2012Natur.483..163G. doi:10.1038/483163a. PMC   4090763 . PMID   22398551.
  45. Bae, Chilman; Sachs, Frederick; Gottlieb, Philip A. (2011). "The Mechanosensitive Ion Channel Piezo1 Is Inhibited by the Peptide GsMTx4". Biochemistry. 50 (29): 6295–300. doi:10.1021/bi200770q. PMC   3169095 . PMID   21696149.
  46. Dedman, Alexandra; Sharif-Naeini, Reza; Folgering, Joost H. A.; Duprat, Fabrice; Patel, Amanda; Honoré, Eric (2008). "The mechano-gated K2P channel TREK-1". European Biophysics Journal. 38 (3): 293–303. doi:10.1007/s00249-008-0318-8. PMID   18369610. S2CID   28802245.
  47. Sackin, H. (1995). "Mechanosensitive channels". Annu. Rev. Physiol. 57: 333–53. doi:10.1146/annurev.ph.57.030195.002001. PMID   7539988.
  48. Sukharev SI, Martinac B, Arshavsky VY, Kung C (July 1993). "Two types of mechanosensitive channels in the Escherichia coli cell envelope: solubilization and functional reconstitution". Biophys. J. 65 (1): 177–83. Bibcode:1993BpJ....65..177S. doi:10.1016/S0006-3495(93)81044-0. PMC   1225713 . PMID   7690260.
  49. Haswell ES, Phillips R, Rees DC (October 2011). "Mechanosensitive channels: what can they do and how do they do it?". Structure. 19 (10): 1356–69. doi:10.1016/j.str.2011.09.005. PMC   3203646 . PMID   22000509.
  50. Ernstrom GG, Chalfie M (2002). "Genetics of sensory mechanotransduction". Annu. Rev. Genet. 36: 411–53. doi:10.1146/annurev.genet.36.061802.101708. PMID   12429699.
  51. García-Añoveros J, Corey DP (May 1996). "Touch at the molecular level. Mechanosensation". Curr. Biol. 6 (5): 541–3. doi: 10.1016/S0960-9822(02)00537-7 . PMID   8805263.
  52. 1 2 Purves, Dale. (2004). Neuroscience. Sunderland, Mass.: Sinauer Associates. pp. 207–209. ISBN   978-0-87893-725-7.
  53. 1 2 3 4 5 6 7 8 9 10 11 Del Valle ME, Cobo T, Cobo JL, Vega JA (August 2012). "Mechanosensory neurons, cutaneous mechanoreceptors, and putative mechanoproteins". Microsc. Res. Tech. 75 (8): 1033–43. doi:10.1002/jemt.22028. PMID   22461425. S2CID   206068242.
  54. 1 2 3 4 Patel A, Sharif-Naeini R, Folgering JR, Bichet D, Duprat F, Honoré E (August 2010). "Canonical TRP channels and mechanotransduction: from physiology to disease states". Pflügers Arch. 460 (3): 571–81. doi:10.1007/s00424-010-0847-8. PMID   20490539. S2CID   22542282.
  55. 1 2 3 4 5 6 7 8 9 López-Larrea, Carlos (2011). Sensing in Nature. New York: Springer Science+Business Media. ISBN   978-1-4614-1703-3.
  56. 1 2 3 4 5 6 7 Yin J, Kuebler WM (2010). "Mechanotransduction by TRP channels: general concepts and specific role in the vasculature". Cell Biochem Biophys. 56 (1): 1–18. doi:10.1007/s12013-009-9067-2. PMID   19842065. S2CID   12154460.
  57. 1 2 3 Martinac B (2011). "Bacterial mechanosensitive channels as a paradigm for mechanosensory transduction". Cell. Physiol. Biochem. 28 (6): 1051–60. doi: 10.1159/000335842 . PMID   22178995.
  58. Peyronnet R, Nerbonne JM, Kohl P (2016). "Cardiac mechano-gated ion channels and arrhythmias". Circ. Res. 118 (2): 311–29. doi:10.1161/CIRCRESAHA.115.305043. PMC   4742365 . PMID   26838316.
  59. Quinn TA, Kohl P (2021). "Cardiac Mechano-Electric Coupling: Acute Effects of Mechanical Stimulation on Heart Rate and Rhythm". Physiol. Rev. 101 (1): 37–92. doi: 10.1152/physrev.00036.2019 . PMID   32380895.
  60. 1 2 3 4 5 6 7 8 Sachs F (2010). "Stretch-activated ion channels: what are they?". Physiology. 25 (1): 50–6. doi:10.1152/physiol.00042.2009. PMC   2924431 . PMID   20134028.
  61. 1 2 Bianchi L (December 2007). "Mechanotransduction: touch and feel at the molecular level as modeled in Caenorhabditis elegans". Mol. Neurobiol. 36 (3): 254–71. doi:10.1007/s12035-007-8009-5. PMID   17955200. S2CID   6474334.
  62. Formigli L, Meacci E, Sassoli C, Squecco R, Nosi D, Chellini F, Naro F, Francini F, Zecchi-Orlandini S (May 2007). "Cytoskeleton/stretch-activated ion channel interaction regulates myogenic differentiation of skeletal myoblasts". J. Cell. Physiol. 211 (2): 296–306. doi:10.1002/jcp.20936. PMID   17295211. S2CID   2800864.
  63. Zhao Y, Yamoah EN, Gillespie PG (December 1996). "Regeneration of broken tip links and restoration of mechanical transduction in hair cells". Proc. Natl. Acad. Sci. U.S.A. 93 (26): 15469–74. Bibcode:1996PNAS...9315469Z. doi: 10.1073/pnas.93.26.15469 . PMC   26428 . PMID   8986835.
  64. Bell J, Bolanowski S, Holmes MH (January 1994). "The structure and function of Pacinian corpuscles: a review". Prog. Neurobiol. 42 (1): 79–128. doi:10.1016/0301-0082(94)90022-1. PMID   7480788. S2CID   45410718.
  65. Tay A, Dino DC (January 17, 2017). "Magnetic Nanoparticle-Based Mechanical Stimulation for Restoration of Mechano-Sensitive Ion Channel Equilibrium in Neural Networks". Nano Letters. 17 (2): 886–892. doi:10.1021/acs.nanolett.6b04200. PMID   28094958..
  66. 1 2 Lumpkin EA, Caterina MJ (February 2007). "Mechanisms of sensory transduction in the skin". Nature. 445 (7130): 858–65. Bibcode:2007Natur.445..858L. doi:10.1038/nature05662. PMID   17314972. S2CID   4391105.
  67. 1 2 3 Lumpkin, Ellen A.; Caterina, Michael J. (2006). "Mechanisms of sensory transduction in the skin". Nature. 445 (7130): 858–865. Bibcode:2007Natur.445..858L. doi:10.1038/nature05662. PMID   17314972. S2CID   4391105.
  68. Markin, V.S.; Martinac, B. (1991). "Mechanosensitive ion channels as reporters of bilayer expansion. A theoretical model". Biophys. J. 60 (5): 1120–1127. Bibcode:1991BpJ....60.1120M. doi:10.1016/S0006-3495(91)82147-6. PMC   1260167 . PMID   1722115.
  69. 1 2 Perozo, E.; Cortes, D. M.; Sompornpisut, P.; Kloda, A.; Martinac, B. (2002). "Structure of MscL and the gating mechanism of mechanosensitive channels". Nature. 418 (6901): 942–8. Bibcode:2002Natur.418..942P. doi:10.1038/nature00992. PMID   12198539. S2CID   4350910.
  70. Hamill, O.P.; McBride, Jr (1997). "Induced membrane hypo /hyper mechanosensitivity A limitation of patch-clamp recording". Annu. Rev. Physiol. 59: 621–631. doi:10.1146/annurev.physiol.59.1.621. PMID   9074780.
  71. Martinac B, Buechner M, Delcour AH, Adler J, Kung C (April 1987). "Pressure-sensitive ion channel in Escherichia coli". Proc. Natl. Acad. Sci. U.S.A. 84 (8): 2297–301. Bibcode:1987PNAS...84.2297M. doi: 10.1073/pnas.84.8.2297 . PMC   304637 . PMID   2436228.
  72. Perozo, E.; Rees, D.C. (2003). "Structure and mechanism in prokaryotic mecahnosensitive channels". Current Opinion in Structural Biology. 13 (4): 432–442. doi:10.1016/S0959-440X(03)00106-4. PMID   12948773.
  73. Levina, N.; Totemeyer, S.; Stokes, N. R.; Louis, P.; Jones, M. A.; Booth, I. R. (1999). "Protection of Escherichia coli cells against extreme turgor by activation of MscS and MscL mechanosensitivechannels: Identification of genes required for MscS activity". The EMBO Journal. 18 (7): 1730–1737. doi:10.1093/emboj/18.7.1730. PMC   1171259 . PMID   10202137.
  74. Bass, R. B.; Strop, P.; Barclay, M.; Rees, D. (2002). "Crystal structure of Escherichia coli MscS, a voltage-modulated and mechanosensitive channel" (PDF). Science. 298 (5598): 1582–1587. Bibcode:2002Sci...298.1582B. doi:10.1126/science.1077945. PMID   12446901. S2CID   15945269.
  75. Pivetti, C. D.; Yen, M. R.; Miller, S.; Busch, W.; Tseng, Y.; Booth, I. R.; Saier, MH (2003). "Two families of mechanosensitive channel proteins". Microbiol. Mol. Biol. Rev. 67 (1): 66–85. doi:10.1128/MMBR.67.1.66-85.2003. PMC   150521 . PMID   12626684.
  76. Vasquez, V.; Sotomayor, M.; Cordero-Morales, J.; Shulten, K.; Perozo, E. (2008). "A Structural mechanism for MscS gating lipid channels in bilayer". Science. 321 (5893): 1210–14. Bibcode:2008Sci...321.1210V. doi:10.1126/science.1159674. PMC   2897165 . PMID   18755978.
  77. Bezanilla, F.; Perozo, E. (2002). "Force and voltage sensors in one structure". Science. 298 (5598): 1562–1563. doi:10.1126/science.1079369. PMID   12446894. S2CID   118927744.
  78. Sukharev, S. I.; Blount, P.; Martinac, B.; Kung, C. (1997). "MECHANOSENSITIVE CHANNELS OF ESCHERICHIA COLI: The MscL Gene, Protein, and Activities". Annu. Rev. Physiol. 59: 633–57. doi:10.1146/annurev.physiol.59.1.633. PMID   9074781.
  79. Sukharev, S. I.; Blount, P.; Martinac, B.; Blattner, F. R.; Kung, C. (1994). "A large mechanosensitive channel in E. coli encoded by MscL alone". Nature. 368 (6468): 265–268. Bibcode:1994Natur.368..265S. doi:10.1038/368265a0. PMID   7511799. S2CID   4274754.
  80. Chang, G.; Spencer, R.; Barclay, R.; Lee, A.; Barclay, M.; Rees, C. (1998). "Structure of the MscL homologue from Mycobacterium tuberculosis: a gated mechanosensitive ion channel". Science. 282 (5397): 2220–2226. Bibcode:1998Sci...282.2220C. doi: 10.1126/science.282.5397.2220 . PMID   9856938.
  81. Blount, P; Sukharev, SI; Moe, PC; Schroeder, MJ; Guy, HR; Kung, C. (1996). "Membrane topology and multimeric structure of a mechanosensitive channel protein". The EMBO Journal. 15 (18): 4798–4805. doi:10.1002/j.1460-2075.1996.tb00860.x. PMC   452216 . PMID   8890153.
  82. Arkin IT, Sukharev SI, Blount P, Kung C, Brünger AT (February 1998). "Helicity, membrane incorporation, orientation and thermal stability of the large conductance mechanosensitive ion channel from E. coli". Biochim. Biophys. Acta. 1369 (1): 131–40. doi: 10.1016/S0005-2736(97)00219-8 . PMID   9528681.
  83. Sukharev, S.; Betanzos, M.; Chiang, C.S.; Guy, H.R. (2001). "The gating mechanism of the large mechanosensitive channel MscL". Nature. 409 (6821): 720–724. Bibcode:2001Natur.409..720S. doi:10.1038/35055559. PMID   11217861. S2CID   4337519.
  84. 1 2 Perozo, E.; Cortes, D. M.; Sompornpisut, P.; Kloda, A.; Martinac, B. (2002). "Open channel structure of MscL and the gating mechanism of mechanosensitive channels". Nature. 418 (6901): 942–948. Bibcode:2002Natur.418..942P. doi:10.1038/nature00992. PMID   12198539. S2CID   4350910.
  85. Wiggins, P; Phillips, R (2004). "Analytic models for mechanotransduction: Gating a mechanosensitive channel". Proc Natl Acad Sci U S A. 101 (12): 4071–6. arXiv: q-bio/0311010 . Bibcode:2004PNAS..101.4071W. doi: 10.1073/pnas.0307804101 . PMC   384697 . PMID   15024097.
  86. Wiggins, P; Phillips, R (2005). "Membrane-protein interactions in mechanosensitivechannels". Biophys J. 88 (2): 880–902. arXiv: q-bio/0406021 . Bibcode:2005BpJ....88..880W. doi:10.1529/biophysj.104.047431. PMC   1305162 . PMID   15542561.
  87. Coste B, Mathur J, Schmidt M, Earley TJ, Ranade S, Petrus MJ, Dubin AE, Patapoutian A (October 2010). "Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels". Science. 330 (6000): 55–60. Bibcode:2010Sci...330...55C. doi:10.1126/science.1193270. PMC   3062430 . PMID   20813920.
  88. Zarychanski R, Schulz VP, Houston BL, Maksimova Y, Houston DS, Smith B, Rinehart J, Gallagher PG (August 2012). "Mutations in the mechanotransduction protein PIEZO1 are associated with hereditary xerocytosis". Blood. 120 (9): 1908–15. doi:10.1182/blood-2012-04-422253. PMC   3448561 . PMID   22529292.
  89. Coste B, Houge G, Murray MF, Stitziel N, Bandell M, Giovanni MA, Philippakis A, Hoischen A, Riemer G, Steen U, Steen VM, Mathur J, Cox J, Lebo M, Rehm H, Weiss ST, Wood JN, Maas RL, Sunyaev SR, Patapoutian A (March 2013). "Gain-of-function mutations in the mechanically activated ion channel PIEZO2 cause a subtype of Distal Arthrogryposis". Proc. Natl. Acad. Sci. U.S.A. 110 (12): 4667–72. Bibcode:2013PNAS..110.4667C. doi: 10.1073/pnas.1221400110 . PMC   3607045 . PMID   23487782.
  90. Engler, A.; Shamik, S.; Sweeney, L.; Disher, D. (2006). "Matrix Elasticity Directs Stem Cell Lineage Specification". Cell. 126 (4): 677–689. doi: 10.1016/j.cell.2006.06.044 . PMID   16923388.
  91. Hamill, O.P.; Martinac, B. (2001). "Molecular basis of mechanotransduction in living cells". Physiol. Rev. 81 (2): 685–740. doi:10.1152/physrev.2001.81.2.685. PMID   11274342. S2CID   1877143.
  92. Nguyen, T.; Clare, B.; Martinac, B.; Martinac, Boris (2005). "The effects of parabens on the mechanosensitive channels". Eur. Biophys. J. 34 (5): 389–396. doi:10.1007/s00249-005-0468-x. PMID   15770478. S2CID   45029899.
  93. Guharay F, Sachs F (July 1984). "Stretch-activated single ion channel currents in tissue-cultured embryonic chick skeletal muscle". J. Physiol. 352: 685–701. doi:10.1113/jphysiol.1984.sp015317. PMC   1193237 . PMID   6086918.
  94. Tang, Y.; Cao, G.; Chen, X.; et al. (2006). "A finite element framework for studying the mechanical response of macromolecules: application to the gating of the mechanosensitive channel MscL". Biophys J. 91 (4): 1248–63. Bibcode:2006BpJ....91.1248T. doi:10.1529/biophysj.106.085985. PMC   1518658 . PMID   16731564.
  95. 1 2 Patel A, Sharif-Naeini R, Folgering JR, Bichet D, Duprat F, Honoré E (2010). "Canonical TRP channels and mechanotransduction: from physiology to disease states". Pflügers Arch. 460 (3): 571–81. doi:10.1007/s00424-010-0847-8. PMID   20490539. S2CID   22542282.
  96. Maingret F, Fosset M, Lesage F, Lazdunski M, Honoré E (January 1999). "TRAAK is a mammalian neuronal mechano-gated K+ channel". J. Biol. Chem. 274 (3): 1381–7. doi: 10.1074/jbc.274.3.1381 . PMID   9880510.
  97. Patel AJ, Honoré E, Maingret F, Lesage F, Fink M, Duprat F, Lazdunski M (August 1998). "A mammalian two pore domain mechano-gated S-like K+ channel". EMBO J. 17 (15): 4283–90. doi:10.1093/emboj/17.15.4283. PMC   1170762 . PMID   9687497.
  98. Nagasawa M, Kanzaki M, Iino Y, Morishita Y, Kojima I (2001). "Identification of a novel chloride channel expressed in the endoplasmic reticulum, golgi apparatus, and nucleus". J. Biol. Chem. 276 (23): 20413–20418. doi: 10.1074/jbc.M100366200 . PMID   11279057.
  99. Ozeki-Miyawaki C, Moriya Y, Tatsumi H, Iida H, Sokabe M (2005). "Identification of functional domains of Mid1, a stretch-activated channel component, necessary for localization to the plasma membrane and Ca2+ permeation". Exp. Cell Res. 311 (1): 84–95. doi:10.1016/j.yexcr.2005.08.014. PMID   16202999.

The following is not referenced in the article, and/or is in conflict with Engler, A. et al., 2006: