A possible unifying principle for mechanosensation Ching Kung1 Of Aristotle's five senses, we know that sight, smell and much of taste are initiated by ligands binding to G-proteincoupled receptors; however, the mechanical sensations of touch and hearing remain without a clear understanding of their molecular basis. Recently, the relevant force-transducing molecules--the mechanosensitive ion channels--have been identified. Such channel proteins purified from bacteria sense forces from the lipid bilayer in the absence of other proteins. Recent evidence has shown that lipids are also intimately involved in opening and closing the mechanosensitive channels of fungal, plant and animal species. A ll creatures have mechanical senses: insects hear, and, when touched, worms twitch and sea anemones contract. Touching the front of a unicellular paramecium makes it swim backward; touching its posterior makes it spurt forward. Plant roots and shoots respond to gravity (gravitropism), and stems proportion the growth in their height, versus the growth in their girth, based on the amount of jostling by wind and rain (thigmomorphogenesis). Besides the ear and the skin, animals have many other mechanosensors; for example, the circumventricular organs (for determining systemic osmolarity), baroreceptors (blood pressure), spindle receptors (muscle stretch), proprioceptors (limb positions) and so on. Our bones measure stress during periods of growth or regeneration, and our tongues sense texture and size so that we don't swallow sand or stones. Mechanical senses differ from other senses. The molecular bases of sensing odorants, hormones, neurotransmitters and other dissolved ligands (solutes) are well understood: the lock-and-key binding of each ligand to the specific binding pocket of its specific receptor on the plasma membrane. Much less is known, however, about the molecules that sense forces such as osmotic force, thirst, touch, vibration and texture. Many membranes are equipped with mechanosensitive (MS) ion channels that respond to turgor in proportion to the surrounding concentration of water (the solvent). Such channel proteins have been cloned and crystallized from bacteria. Examination of these proteins by genetic, electric, chemical and physical means has found that they are able to directly detect and respond to forces from the lipid bilayer. The study of MS channels in plants and animals lags behind, partly because their anatomical complexities resist reductionistic approaches. Nonetheless, recent findings indicate the involvement of lipids in the gating of MS channels of the worm, the fly, the frog oocyte and in mammals. The same types of channel proteins--for example, transient receptorpotential (TRP) proteins--are now found to sense vibration, touch and osmotic membrane stretch. The possibility is emerging that force detection ultimately occurs at the channel­lipid interface. Displacement at the interface, either by deforming the bilayer or by pulling the channel or the bilayer with a tether, provides the energetic drive for the channel to reach the open conformation. This Review covers many of the principal advances in this field, and draws on widely diverse research projects: from hair cells to bacteria. Traditional compartmentalization of research disciplines, such as the segregation of micro- and neuro-biology, often hampers dialogue among projects. Yet, life's basic machineries--such as DNA replication, transcription, translation, the Krebs cycle, electron transport, and now ion filtration and voltage gating--are found to be universal, even though they were originally based, in large measures, on experiments done in microbial systems. This Review asks whether there is a common physicochemical basis that determines how channel proteins sense force, which may serve to unite the varied biological manifestations. Even though touch, hearing and osmosis are seemingly disparate fields of research, they all deal with a single physical parameter--force. Regardless of frequency or duration, a dyne is a dyne. MS channels allow bacteria to withstand rain Life is largely aqueous chemistry--we went to Mars looking for evidence of water and, by inference, life. Cells comprise ,80% water, and de- or over-hydration can be lethal. Osmotic force, a measure of water content, is therefore fundamental to cell survival. Even though water is 55.6 M, it only requires a 10 milli-osmolar difference across a barrier to produce an osmotic pressure of ,180 mm Hg (,2.5 105 dyn cm2 2 ), which generates a ,12 dyn cm2 1 stretch force on the surface of a 2-mm-diameter sphere. Surprisingly, the proteins in bacteria that directly gauge such stretch forces were only discovered in the last decade, and only by serendipity. When a cell is caught in the rain, or following laboratory dilution, the inward diffusion of water through the lipid bilayer--water channels are not needed--generates a huge turgor that can rise to hundreds of atmospheres (108 dyn cm2 2 ): far beyond what the cell envelope can withstand. It is known, from work first carried out in the 1950s, that Escherichia coli bacteria release their osmolites (solutes) upon an osmotic down-shock (Fig. 1a). For instance, a 1-in-100 aqueous dilution of the culture drains the cellular 14 C-labelled proline pool by .95% (ref. 1). Tracing other osmolites, such as K , lactose and ATP, gave similar results2 . However, osmolite-depleted bacteria retain their macromolecules, and begin protein synthesis within minutes upon return to normal medium1 . The identity of the `emergency valves' for solute release remained elusive for almost half a century. Though MscL and MscS (mechanosensitive channel of large or small unitary conductance) were postulated to be these valves when their electrical activities were discovered in 1987 (ref. 3), this role was not proved until 1999 when Booth and co-workers4 showed that DmscL REVIEWS 1 Laboratory of Molecular Biology and Department of Genetics, University of Wisconsin, 1525 Linden Drive, Madison, Wisconsin 53706, USA. Vol 436|4 August 2005|doi:10.1038/nature03896 647 2005 Nature Publishing Group DmscS double mutants lyse even upon a rather mild down-shock. MscL and MscS have redundant functions and single mutants have no clear phenotype, so they were not discovered in the extensive and intensive genetic dissection of E. coli physiology in the last century, but were, in fact, identified in a patch-clamp survey (discussed below). MS channels are just one--albeit the quickest one--of a bacterium's many defences against de- or over-hydration. The slower types of these defences are those that are transcriptionally controlled. Prokaryotic MS channels MscL and MscS were encountered in the first electrophysiological survey of bacterial membranes, conducted by Martinac et al.3 . Voltage-controlled (-clamped) patches of E. coli membranes produced giant steps in unitary current when the patches were subjected to suction3 , or when the solution bathing the patch was diluted5 . These MS-channel activities remain following reconstitution into artificial liposomes after membrane dissolution or protein purifi- cation6 (Fig. 1b). The non-selective unitary conductances of MscL (,3 nS) and MscS (,1 nS) are 10­1,000-fold greater than those of the more selective channels commonly studied. Such large signals and unlimited bacterial material allowed the tracing of the channel activity in chromatographic fractions to a single protein, which, in turn, led to the discovery of the mscL gene7,8 . The corresponding small protein has two transmembrane (TM) helices, M1 and M2 (refs 7, 9; Fig. 2, left). Chang et al. resolved the crystal structure of the Mycobacterium tuberculosis MscL homologue as a homopentamer with the five M1 domains converging to close the pore at the cytoplasmic side10 (Fig. 2, centre). In 2001, computational modelling and cross-linking experiments led Sukharev et al.11 to a model in which all TM helices recline and rotate, like the iris of a camera, to open a large pore of some 30 A° in diameter (Fig. 2, right). The following year, through observations with electron paramagnetic resonance spectroscopy after site-directed probe attachment, Martinac and co-workers12,13 concurred with the main premise of this model. The cloning4 and crystallization14 of the E. coli MscS protein showed that it is a homoheptamer of three-TM subunits with seven converging M3 domains: entirely different from MscL. Genetic, biochemical, biophysical, simulation and other studies on MscL and MscS are periodically reviewed15­18 . Forces from lipids gate prokaryotic MS channels. Exercising ultimate reductionism, mechanosensitivity was found to be intact when pure MscL protein was placed in bilayers of one12,13 or two7 defined lipids (Fig. 1b). Here, the stretch force detected by the protein must come from the lipids themselves, since there are no other components that could contribute. MscL follows a Boltzmann distribution where the mechanical energy partitions the MscL protein between its closed and open conformations, with the midpoint tension (50% of the channel being open) at ,12 dyn cm21 (ref. 19)--a sensitivity presumably tuned for its biological role (Fig. 1a). (The threshold tension of MscL is much lower; wild-type MscS (ref. 20) and MscL gain-of-function mutants21,22 have even lower thresholds.) MscL functions in bilayers made of ordinary lipids with charged or uncharged heads, saturated or unsaturated tails, and in various mixtures. Shortening the length of the fatty acid chain from 20, to 18, to 16 carbons reduces the energy barrier between the closed and the open state, but does not trigger spontaneous channel opening12,13 . The prevailing model of the behaviour of MscL and MscS addresses the forces within the lipid bilayer itself. Unlike the aqueous solution, the bilayer is highly anisotropic: having very different physical properties at different depths. The free-energy reduction in ordering waters and lipids at the interface is reflected in a large surface tension between the lipids' polar head groups and the nonpolar tails. However, pressures nearby serve to balance this tension, allowing the bilayer to self-assemble into a stable structure. The force profile of pure lipid bilayers has been calculated by Cantor23,24 (Fig. 3a) and examined with molecular-dynamics simulation25,26 . (How the protein­lipid interaction may distort the profile at the interface is unclear, though it is being analysed27 .) The amplitudes of these forces are in the order of hundreds of dyne cm2 1 (ref. 25); much larger than the lytic tension (tens of dyne cm2 1 ) of the bilayer. Any protein embedded in the bilayer is subject to these strong, localized push or pull forces. Altering the force profile by membrane stretch, or by channel or lipid displacement through a tether (discussed below), may make it energetically more favourable for the channel protein to assume a new conformation; for example, the open state (Fig. 3b). Among the molecular-dynamics simulations carried out by several groups, Gullingsrud and Schulten26 directed forces to residues within the MscL protein at the level of the lipid's glycerol backbone between the head groups and the tails, where tension is maximal (yellow arrows, Fig. 2). Such a simulation can reveal steric clashes or structural disintegration, which should not happen when a channel opens. They traced the positions of the 111,079 atoms of 1 MscL protein, 365 lipids and 22,308 waters, and found that MscL indeed opens on a 10-ns timescale in an iris-like manner, similar to the original model11­13 . Besides external forces, the composition of the lipid bilayer itself can affect its internal forces. Adding chemically unrelated cationic Figure 1 | Bacterial channels function as emergency release valves in vivo, and the mechanosensitivity of pure MscL channel protein in vitro. a, An E. coli cell in a normal environment (left) and in the rain (or upon dilution in the laboratory, right). A bacterium (shown as a rod), having adjusted its cytoplasm to the relatively high osmolarity of the surrounding milieu (shown in dark red, the red dots being solutes, not water), is confronted with a sudden dilution of its environment upon the onset of rain (light red). Entry of water (not shown) through the lipid bilayer swells the bacterium (now oval-shaped) and stretches open the MS channels to jettison solutes (red puffs), enabling it to reach a new equilibrium and escaping osmolysis (and returns to being rod-shaped). b, Purified MscL protein is reconstituted into multilamellar liposomes after replacing the detergent with lipids. Induced liposome blisters can be sampled with a patch-clamp pipette. A suction applied to the pipette (broad open arrow) creates tension (small filled arrows) in the membrane patch and activates MscL proteins. The increase in the number of channel openings in a patch (shown as unitary-conductance steps at the marked levels) is evident when the suction applied to such a patch of lipid bilayer increases from 30 mm Hg to 40 mm Hg (4 104 dyn cm2 2 to 5.3 104 dyn cm22 ) (modified from ref. 15). REVIEWS NATURE|Vol 436|4 August 2005 648 2005 Nature Publishing Group amphipaths to the membrane of a red blood cell causes it to form cups, while adding anionic amphipaths causes it to form bulges (to crenate)28 . Regardless of the cause, changes in the bilayer's geometry will distort the force profile contained within it. Indeed, these agents have been shown to reversibly activate MscL and MscS. The potency of the amphipaths to activate such channels is proportional to their lipid solubility29 , and they are effective only when added to one monolayer but not both13 . The shapes of the added lipids are important, as evident from the behaviour of gramicidin A upon bilayer modifications uncovered by Andersen and co-workers30 . The usual bilayer-forming phospholipids can be approximated as rods (Fig. 4, shown in red), and the micelle-forming lysophospholipids (with a single fatty-acid chain) as cones (Fig. 4, blue). Polyunsaturated fatty acids (PUFAs), such as arachidonic acid (AA, a precursor of prostaglandins), can be regarded as inverted cones with smaller heads than tails (Fig. 4, green). Cones, or inverted cones, differentially entered into one monolayer can change the local curvature and the internal force profile, redistributing the tension between the two leaflets (Fig. 3a). Indeed, the addition of lysophosphatidylcholine triggers the opening of MscS (ref. 13). Structurally diverse anaesthetics, which are all lipid soluble, have been theorized to change the bilayer force profile31 . Indeed, experiments have shown that procaine and tetracaine can activate MscS (ref. 29). Eukaryotic MS channels Plants (for example, Arabidopsis thaliana) have clear homologues of MscS. Most animal cell membranes present MS conductances, but only a few have been traced to known gene products. Although these differ from MscS and MscL in sequence, some of their properties are quite similar. Patel and co-workers found that the polymodal K channel of mammals, TREK-1 (two-pore domain weak inwardrectifying (TWIK)-related K channel), can be activated by both force and osmolarity. Similar to MscS and MscL, it is also activated by the bulge-forming amphipathic chemicals such as trinitrophenol (called crenaters), but inhibited by the cup-forming amphipaths such as chlorpromazine32,33 . The cone-shaped lysophosphatidylcholine activates it, and the exaggerated cone lysophosphatidylinositol activates even more effectively32,34 . Structurally diverse anaesthetics such as chloroform, halothane, isoflurane and diethyl-ether also activate TREK-1 (ref. 35). Negatively charged lipids such as PIP2 (phosphatidylinositol 4,5-bisphosphate) or phosphatidic acid, when presented to the membrane inner leaflet, also activate TREK-1 (ref. 36), and the charged cone lysophosphatidic acid strongly activates it37 . MS and some other types of channels are inhibited by the small lanthanide Gd3 (refs 38, 39). The mechanisms of this inhibition are complex and include its interaction with the lipid bilayer39,40 . Sachs and others found an amphipathic 34 L-amino-acid peptide41 in a tarantula venom to inhibit the MS conductance in cultured mammalian cells42 . They also found the synthetic enantiomer of 34 D-amino-acids to be just as effective43 , and concluded that the peptide could not have bound to a channel protein lock-and-keylike, but entered the bilayer to affect the channel's surrounding environment. Eukaryotic cells have extensive cytoskeletons near the bilayer, which are often assumed to be the transmitters of force. One needs to examine this assumption carefully. Hamill et al. examined the MS channels in the complex cell surface of Xenopus oocytes44 . From this surface, they induced blebs that had little or no cytoskeletal elements and continued to encounter MS-channel activities in the bleb membranes16,44 , and it was from such membranes they now traced the channel activities to TRPC1 (transient receptor-potential canonical 1; ref. 45; discussed below). The cortical cytoskeleton network often folds a large excess of membrane bilayer into microvilli or cavaeolae. This network is far more extendable than the bilayer itself, allowing cells to swell without increasing the total bilayer area or tension, which explains why MS currents sometimes cannot be found in whole cells but only in excised patches, where the cytoskeleton is lost46,47 . Animal sensory cells often have organized microtubules such as the ciliary axoneme (see below). There are also specialized microtubular processes along the length of the long touch-sensing cells in Caenorhabditis elegans studied by Chalfie and co-workers48,49 . The touch-insensitive (loss-of-function) or touch-cell degeneration (gain-of-function)48­50 worm phenotypes led elegantly to the identification of a series of mechanosensitivity genes (the mec genes) Figure 2 | Opening MscL in E. coli. Helical segments (S1, S2, S3) and transmembane helices (M1, M2) in one MscL subunit, as deduced from sequence7 and other analyses9 (left). Side (upper centre) and top (lower centre) views of the closed channel backbone structure of the E. coli MscL protein, by analogy to the crystal structure of the M. tuberculosis MscL homologue10 . The open structure deduced from both modelling and experimentation11 (right). Unlike MthK, the prokaryotic K channel that is equipped with a second constriction (the K filter), MscL is like the acetylcholine receptor/channel, in which the open gate doubles as the filter. Here the opening is huge (,30 A° in diameter): befitting its ability to release solutes indiscriminately (as shown in Fig. 1a). The work to increase the area under tension constitutes the free energy difference that partitions the open and closed states. (Modified from ref. 11.) NATURE|Vol 436|4 August 2005 REVIEWS 649 2005 Nature Publishing Group including mec-4 and mec-10. These genes encode channel subunits akin to the epithelial Na channels, and conduct channel currents in frog oocytes51,52 . The channels, and their associated components, form punctate clusters that are evenly spaced along the worm's microtubular processes. Among the MECs are an extracellular matrix protein (MEC-1) and special tubulins (MEC-7, MEC-12). Previous models described a trans-cellular complex, similar to the original trapdoor model for the vertebrate hair cell (now modified, see below), in which the displacement of the matrix, resisted by the microtubules, gates the channels in between48,49 . Recent analyses showed that MEC-1, and at least two other gene products in the matrix, are needed to form the punctate clusters. However, removal of the special microtubules through mutation have only limited or no effect on the structure or the function of these clusters53,54 . These observations, and the fact that the transduction current is reduced but not abolished in mec-7 mutants55 , have now questioned whether the MS channel is directly gated by tethering to microtubules53 . Tethering to rigid elements, even if true, does not necessarily imply force transmission. Current cell biology teaches us that meaningful protein­protein contacts, whether ephemeral or long-lasting, are the norm. For example, the signalling machinery of the Drosophila photoreceptor, not known for its mechanosensitivity properties, is comprised of rhodopsin, G protein, enzymes and channels, all tied together into a "transducisome"56 or "signalplex"57 tethered to an actin cytoskeleton, which serves to deploy and station the complex near the cell surface. What trips TRP channels? Several MS channels making recent news belong to the TRP super- family45,58,59 , the founding member of which was identified in a near-blind Drosophila with a transient receptor-potential in its electroretinogram60 . Forward genetics starting from mechanoinsensitive phenotypes led us to these TRP channels without preconceived bias. This has happened six times recently--in the worm61,62 , fly63,64 , mouse65 and human66 --and cannot be a coincidence. At least seven homologues of the TRPs identified in this way then became candidates in the ensuing investigations and have now also been tied to mechanosensation. TRPV4 (transient receptor-potential vanilloid-4; previously called the vanilloid receptor-related osmotically activated channel, or VROAC), the most studied mammalian MS TRP channel, is found in many tissues including the circumventricular organs of the central nervous system and inner-ear hair cells. The heterologously expressed whole-cell current through TRPV4 can be activated by cell inflation with pressure67 , by mechanical shear force from bath flow68 , or by mild hypo-osmotic challenges69,70 . Deleting the three amino-terminal ankyrin repeats does not significantly impair TRPV4's response to hypo-osmolarity69 . The Bargmann laboratory found the first MS TRPV through osm-9-mutant worms, which fail to recoil from an osmolar solution or a nose touch71 . Worms with normal OSM-9 channels, but deficient in the synthesis of a set of 20-carbon PUFAs (including AA, Fig. 4), showed similar deficits in behaviour that could be restored by a dietary supplement of PUFAs. One such PUFA acts as an irritant that appears to directly activate the channel72 . The rat TRPV4 is only 24% identical to the worm's OSM-9. Nonetheless, Liedtke et al. found that a trpv4 transgene complements the defects of the osm-9-mutant worm, and that the restored behaviours have a threshold and temperature optimum of a warm-blooded animal73 . Furthermore, a version of TRPV4 deleted of both its amino- and carboxy-terminal cytoplasmic domains, the presumed cytoskeleton-binding sites, remained able to complement the phenotype73 . The single-channel current of TRPV4 has received far less scrutiny. Unitary conductances of 310 pS (ref. 69), 60 pS (ref. 74), or 30 and 88 pS (ref. 70) have been variously reported from similar experiments. The only report on attempts to activate TRPV4 unitary conductance by direct patch suction was negative70 . In contrast, a 20 pS MS conductance (in frog Ringer's solution) from oocyte membranes has recently been traced to TRPC1 through liposome reconstitution by Maroto et al.45 , who also showed that heterologously expressed human TRPC1 produces MS unitary currents. Although reconstitution of purified TRPC1 has not yet been reported, this work comes closest to showing that certain TRP channels receive their gating force directly from the lipids. Channels associated with mechanosensation appear in nearly all TRP subfamilies: TRPV, TRPC, TRPA (ankyrin-like), TRPP (polycystin), TRPN (NOMPC, or no mechanoreceptor potential C), TRPY (yeast) and probably TRPML (mucolipins), each having rather different cytoplasmic domains, suggesting that these domains are not Figure 3 | The intrinsic forces in the lipid bilayer, and how applied forces can open MS channels. a, The intrinsic force profile plotted as its direction and magnitude along the depth of the bilayer (left)23 , and a cartoon of a channel protein in section (right), showing how the sharp tension (narrow arrows) near the lipid necks balanced by more diffused pressure nearby (broad arrows) is exerted on the channel­lipid interface (red). b, The forces at the crucial channel­lipid interface (red) will change when the bilayer (green) is stretched or bent (left), or when the channel is displaced from the bilayer through a tether like an elevator (right). It is also possible that the tether, through ancillary proteins, pulls on the lipids surrounding the channel (not shown). In all cases, changes in the force profile at the interface (red) can become the ultimate trigger for the channel's conformational change. Figure 4 | The shape of bilayer components affects its geometry and intrinsic forces. a, Bilayer-forming phospholipids (shown in red), such as phosphatidylcholine (PC), can be approximated as rods. Micelle-forming lysophospholipids (blue), such as lysophosphatidylcholine (LPC) with only one fatty acid chain, can be regarded as cones. Polyunsaturated fatty acids (green), such as arachidonic acid (AA), approximate inverted cones. b, The differential addition of cone-shaped lipids (or other amphipaths, not shown) into one of the monolayers can alter the shape, and therefore the intrinsic forces, of the bilayer. (Modified from ref. 32.) REVIEWS NATURE|Vol 436|4 August 2005 650 2005 Nature Publishing Group the crux of mechanosensitivity. TRP channels are polymodal. Heterologously expressed TRPV4, for example, can also be activated by heat, by a phorbol ester, by anandamide and by AA Fig. 4). The well-known heat-sensing vanilloid receptor, TRPV1, identified in the Julius laboratory75 , is also activated by low pH and endogenous inflammatory ligands. The trpv1 knockout mice and their bladder urothelium have abnormal response to hypo-osmolarity76 . Though a universal switch for different mode of stimuli may be appealing, there is evidence that different stimuli use different pathways to open TRPV4 (ref. 77). Strings attached--the roles of tethers Several TRP channels have now been located within some of the complex MS organs. For example, in the fly chordotonal organ, a matrix material presses against the dendritic cap on top of the sensory cilium, which swells into a dilation about one-third of the way from the tip. Kim, Kernan, and their co-workers found that the only two TRPV channels encoded within the Drosophila genome, NAN (encoded by the Nanchung gene)78 and IAV (encoded by the Inactive gene)79 , apparently form heteromeric channels that transduce vibrations into receptor potentials. NAN and IAV proteins are found on the cilia but not the rest of the neuron, and it is restricted at and below the ciliary dilation, some distance from the tip (Fig. 5a). How the movement of the cap-cell matrix translates into the forces experienced by the IAV-NAN channel in the ciliary membrane that lies below it is unclear. The NAN (ref. 78) and IAV (ref. 79) proteins have been individually expressed in cultured cells, where they confer hypo-osmotically induced whole-cell current and a rise in cytoplasmic Ca2 levels. A parsimonious interpretation would be that the channel experiences the vibration as a stretch along the membrane plane. Whether there are other proteins that tie them to the axoneme, whether such tethers transmit force, and whether there are roles for other channel subunits (for example, NOMPA, or no mechanoreceptor potential A, located at the distal end of the cilium)79 await clarification. The vertebrate hair cell is a clear case in which the gating force is passed on to the MS channel through a tether; though it is debatable whether the tether directly pulls a certain domain from the rest of the channel protein, which is held in place by a resistive force, as originally proposed. Molecular identifications recently enabled great strides in this system. First, TRPN was found in sensory hair cells, and to be required for hearing and balance in zebrafish80 . Then, cadherin 23 was discovered to be a major component of the tip link81,82 , and judged to be too stiff to comprise the gating spring. Recently, Corey et al.83 showed that TRPA1 messenger RNA appears at the appropriate time during hair-cell development, and that TRPA1-expression knock-down curtails transduction in vivo. TRPA1 protein is found at the upper part of the stereocilia, though not just at the very tip. It is also located in the pericuticular zone (perhaps for secretion) and, when present, the entire kinocilium (the true cilium with a microtubular axoneme) (Fig. 5b). With such rapid progress, we are looking forward to the identification of other gene products in this system in the very near future. Meanwhile, fleshing out the biophysical scheme with the molecules so far identified has already led to modifications of the familiar `trapdoor' model of hair-cell mechano-transduction (Fig. 5c, left). Calculation and simulation show that the long N-terminal ankyrin repeats of TRPA1 are compatible with the elastic property of the gating spring83,84 (Fig. 5c, right). (The cartoons in Fig. 5c, and similar representations elsewhere, should not be taken literally as the site(s) and the nature of string attachment, in addition to the number and identity of channel subunits and other elements, are unknown. The upper tether can be attached to the gating springs beneath the membrane instead of the channel body.) In the more thoroughly examined multimeric channels, such as MthK (the K channel of Methanobacterium thermoautotrophicum; ref. 85), MscL (ref. 12) and MscS (ref. 14), the channel gates open like the iris of a camera. It is not yet clear how the one-dimensional movement of a stereocilium leads to the twodimensional opening of the TRP tetramers on the side of the cilium, and whether or how the ciliary membrane may move during transduction. TRPA1 is also expressed in neurons of the dorsal root ganglia, the trigeminal nerve, and photoreceptors, making one wonder for which functions the ankyrin repeats serve in these other locations. A role of ankyrin in assembling a TRP tetramer has been suggested86 . Returning to the main theme of this Review, the vertical movement of a channel by the tip link does not necessarily negate the involvement of the lipid bilayer. It seems possible that such `elevator'like movement would still eventually displace the channel with respect to the bilayer's intrinsic force profile (Fig. 3b, right). The mismatch and asymmetry produced by the displacement can, in fact, Figure 5 | TRP channels in auditory sensory cells. TRP channels have been located in complex auditory sensory cells, even though the mechanism by which ciliary vibrations (arrow pairs) lead to the iris-like opening of the channels on the side of the cilia is not clear. a, The antennal chordotonal organ of Drosophila. CM, cap-cell matrix; DC, dendritic cap; CD, ciliary dilation. Red marks the location of NAN (a TRPV-type channel subunit encoded by the Nanchung gene). (Redrawn from ref. 78.) b, Avertebrate hair cell. St, stereocilia; K, kinocilium; PZ, pericuticular zone. Red marks the location of TRPA1. (Modified from ref. 83.) c, Models of the vertebrate haircell transduction channel. Molecular identifications have transformed the biophysical trapdoor model (left) to one with a TRPA channel and a stiff cadherin-containing tip link (right). The elastic element of transduction is now assigned to the ankyrin repeats in the four (presumably) TRPA subunits83,99 (shown as coils), which are presumed to be attached to cytoskeleton and/or myosin (not shown). This current model is compatible with one in which the displacement of the channel protein, with respect to the lipid bilayer, ultimately triggers the channel conformation change as shown in Fig. 3b, right. However, none of these models should be taken literally since we do not yet know the true composition of the transduction channel(s) and how the various channel components contact each other and the lipids. See the main text for some possible variations. NATURE|Vol 436|4 August 2005 REVIEWS 651 2005 Nature Publishing Group be the final mechano-energetic trigger for the required channelconfiguration change. It is even possible that the tether pulls on the rim of an elastic carrel, say, and passes the force through the lipids that surround the channel (Fig. 3b, left), but this possibility has not been investigated. The `trapdoor' (Fig. 5c, left) is usually interpreted as the separation of individual protein domain(s) from the rest of the channel protein by mechanical work17 . The `elevator' entails displacing the entire channel protein from its normal lipid environments, generating tension around the entire circumference, and leading to an iris-like opening (Fig. 3b, right). Either the trapdoor or the elevator model can accommodate the mechanical resistive and elastic elements traditionally described. It is difficult to imagine TRPA1 being indifferent to the lipids, given its activation by lipid-like compounds. Gillespie and coworkers showed that PIP2 localizes towards the tip of the hair bundle and is required for both the mechanical transduction current and its adaptations87 . The mouse TRPA1 was first reported to be a sensor of noxious cold88 . It is also activated by bradykinin, and the oils of mustard, cinnamon, wintergreen and the like, as well as cannabinoids89,90 . Solute senses versus solvent senses Specialized sensory cilia develop from embryonic primary cilia, and have evolved from motile cilia similar to the ones still found in Paramecium, Chlamydomonas and so on. Can we trace the origin of TRP-based mechanosensation beyond motile protists? An MS TRP channel is found in the vacuolar membrane of yeast91 , where it detects osmotic forces in vitro92 and in vivo93,94 . Because all cells have to deal with osmotic force, it may hold a key to the evolutionary origin(s) of mechanosensation. MscL and MscS proteins are found in most free-living Bacteria and Archaea. Thus, the principle of mechanical gating by forces from the lipid bilayer most likely evolved before the divergence of these two domains of life some 3.5 billion years ago. It makes teleological sense for these devices to have evolved early on. When early cells separate two solutions, and horde solutes into one, a water gradient--and therefore a turgor--must develop on the chemiosmotic partition. This turgor has apparently been used ever since, as all extant cells today still have to be turgid to be in a growing steady state. This turgidity helps to break the bonds in the network of rigid elements (membrane, cell wall, extracellular matrix, cytoskeleton and so forth) so that new material can be inserted. Off steady states--the sudden large rise in turgor at the onset of rain (over-hydration) and the large fall in turgor in prolonged hot sun (dehydration)--are likely to have exerted selective pressure on early cells to evolve mechanosensors such as the ancestors of TRP, MscS and MscL. Once the basic principle of activation by lipid forces is employed, it seems unlikely that nature would abandon the principle in the detection of other forces later on in evolution, even as newer generations of protein types out-compete the old. As reviewed above, many extant TRP channels in animals still seem to respond to osmotic forces exerted on the lipid bilayer even though their relatives are specialized for hearing, balance, touch, or texture. Today, we find a myriad of surface receptors for irritants, odorants, hormones, neurotransmitters and growth factors on cell surfaces. They are not sequence homologues, but they could have originated from one or a few receptors by divergent, as well as convergent, evolution. More importantly, the same physicochemical principle of lock-and-key fitting underlies all ligand sensing. Yet one cannot imagine a protein with a lock-and-key water-binding site that discriminates the milli-osmolar differences that cells do, while water exists in tens of molar. Rapid changes in water concentration must therefore be sensed with a different mechanism. From this perspective, it is not difficult to imagine that a parallel myriad of mechanosensors may have evolved from some simple osmotic-forcesensing device in early membranes95 . Figure 6a shows an `ur-cell' with the two types of receptors for the various solutes (for example, nutrients and wastes) and for the one solvent (water). Figure 6b shows a schematic representation of how different senses might have evolved throughout the 3.5 billion years of evolution. This is not a molecular dendrogram. The branching pattern has no precise meaning except that the senses were few in number in the beginning, and that sensing of nutrients and water is fundamental, ancient and disparate. This diagram emphasizes the distinctness of the two classes of sensations that originated to deal with solutes versus solvent--the two basic ingredients of life's chemistry. Complexities and unknowns The disparity between solute sensing and solvent sensing can help clarify our thinking. Many solute (ligand) receptors do not require a membrane. It is also `crystal clear' (from X-ray diffraction studies) that some MS channels have no specific ligand-binding sites (for example, MscL, MscS). These examples should not lead us to assert that no receptors employ both principles. Given nature's propensity to tinker opportunistically, it is even likely for some receptor proteins to evolve specific ligand-binding pockets that face the two aqueous compartments, as well as a lipid-facing surface that transduces forces. This complexity may explain certain polymodality74 or Figure 6 | The disparate sensing of solutes and solvent. a, A diagram of an imaginary early cell equipped with two types of receptors that are required to sense solutes and solvents--the two ingredients of life's chemistry. The dots in the grey background represent water molecules (the solvent) and the red circles represent solutes (molecules dissolved in water). When a cell accumulates solutes, the internal water concentration is reduced and the tendency of water to enter the cell results in a turgor. Both the lock-and-key type of receptors (red) for different solutes (ligands), as well as the turgor sensors (blue) for water (the solvent), are needed for even an early cell to survive. b, A hypothetical diagram (not to be mistaken for phylogenetic trees) on the grouping of various senses that emphasizes the discrete separations of the lock-and-key type of sensing of the solutes (red) from the force-from-bilayer type of sensing of the solvent (blue). A further description can be found in the text. (Modified from ref. 100.) REVIEWS NATURE|Vol 436|4 August 2005 652 2005 Nature Publishing Group other differences (for example, the TRPV-dependent behaviour of live worms requires certain features of PUFAs but not necessarily AA72 , while heterologously expressed TRPV4 in vitro seems to require an AA metabolite74 ). Complexities can also arise from possible `signalling lipids' that may act at the protein­lipid­water junction. For example, adding negative charges to the inner leaflet with PIP2 (ref. 87) will have local electrostatic effects36,37 as well as secondary effects on lipid distribution and bilayer force profile. In On the Soul (ref. 96), Aristotle argued that there could be no more than five senses, and folded the senses of hot and cold into touch. He might have been on to something. The detection of heat and that of impact seem to convolve in biology. Heat-induced bilayer rearrangement may alter the membrane tension that gates the polymodal TRP channels97 . Teleologically, have they been designed to sum heat and force (for example, into `pain')? Mechanistically, can we separate the force sensor from the heat sensor within the same protein? The truism that everything is mechanical, and heatsensitive, is not helpful. How can a protein­lipid­water complex be designed to be especially sensitive to a rise in thermal agitation? And how can a similar complex become sensitive to its fall? How can some TRPs be constructed to have an unusually high Q10 value (in excess of 10)? If the thermosensor is coupled to a voltage sensor98 , what is the coupling mechanism? Doesn't the temperature sensitivity of a channel­lipid complex depend on the entropy of its hydrophobic interactions with water? Why are the temperature mimics (such as capsaicin, menthol and ginger) oily? Why are the agents that numb the sensation amphipathic? Why is the potency of different general anaesthetics proportional to their solubility in olive oil (the Meyer­Overton rule)? Could the protein­lipid­water junction hold even more secrets? Much work lies ahead if the answers to these questions are to be found; perhaps, just like after a long spell of rain, the floodgates of knowledge are about to be opened. 1. Britten, R. J. & McClure, F. T. The amino acid pool in Escherichia coli. Bacteriol. Rev. 26, 292­-335 (1962). 2. Berrier, C., Coulombe, A., Szabo, I., Zoratti, M. & Ghazi, A. Gadolinium ion inhibits loss of metabolites induced by osmotic shock and large stretchactivated channels in bacteria. Eur. J. Biochem. 206, 559­-565 (1992). 3. Martinac, B., Buechner, M., Delcour, A. H., Adler, J. & Kung, C. Pressuresensitive ion channel in Escherichia coli. Proc. Natl Acad. Sci. USA 84, 2297­-2301 (1987). 4. Levina, N. et al. Protection of Escherichia coli cells against extreme turgor by activation of MscS and MscL mechanosensitive channels: identification of genes required for MscS activity. EMBO J. 18, 1730­-1737 (1999). 5. Martinac, B., Delcour, A. H., Buechner, M., Adler, J. & Kung, C. Advances in Comparative and Environmental Physiology 3­-18 (Springer, Heidelberg, 1992). 6. Delcour, A. H., Martinac, B., Adler, J. & Kung, C. Modified reconstitution method used in patch-clamp studies of Escherichia coli ion channels. Biophys. J. 56, 631­-636 (1989). 7. Sukharev, S. I., Blount, P., Martinac, B., Blattner, F. R. & Kung, C. A largeconductance mechanosensitive channel in E. coli encoded by mscL alone. Nature 368, 265­-268 (1994). 8. Sukharev, S. I., Martinac, B., Blount, P. & Kung, C. Functional reconstitution as an assay for biochemical isolation of channel proteins: application to the molecular identification of a bacterial mechanosensitive channel. Methods: A Companion to Methods in Enzymology 6, 51­-59 (1994). 9. Blount, P. et al. Membrane topology and multimeric structure of a mechanosensitive channel protein of Escherichia coli. EMBO J. 15, 4798­-4805 (1996). 10. Chang, G., Spencer, R. H., Lee, A. T., Barclay, M. T. & Rees, D. C. Structure of the MscL homolog from Mycobacterium tuberculosis: a gated mechanosensitive ion channel. Science 282, 2220­-2226 (1998). 11. Sukharev, S., Betanzos, M., Chiang, C. S. & Guy, H. R. The gating mechanism of the large mechanosensitive channel MscL. Nature 409, 720­-724 (2001). 12. Perozo, E., Cortes, D. M., Sompornpisut, P., Kloda, A. & Martinac, B. Open channel structure of MscL and the gating mechanism of mechanosensitive channels. Nature 418, 942­-948 (2002). 13. Perozo, E., Kloda, A., Cortes, D. M. & Martinac, B. Physical principles underlying the transduction of bilayer deformation forces during mechanosensitive channel gating. Nature Struct. Biol. 9, 696­-703 (2002). 14. Bass, R. B., Strop, P., Barclay, M. & Rees, D. C. Crystal structure of Escherichia coli MscS, a voltage-modulated and mechanosensitive channel. Science 298, 1582­-1587 (2002). 15. Sukharev, S. I., Blount, P., Martinac, B. & Kung, C. Mechanosensitive channels of Escherichia coli: the MscL gene, protein, and activities. Annu. Rev. Physiol. 59, 633­-657 (1997). 16. Hamill, O. P. & Martinac, B. Molecular basis of mechanotransduction in living cells. Physiol. Rev. 81, 685­-740 (2001). 17. Sukharev, S. & Corey, D. P. Mechanosensitive channels: multiplicity of families and gating paradigms. Sci. STKE doi:10.1126/stke.2332004eg7 (2004). 18. Blount, P. Molecular mechanisms of mechanosensation: big lessons from small cells. Neuron 37, 731­-734 (2003). 19. Sukharev, S. I., Sigurdson, W. J., Kung, C. & Sachs, F. Energetic and spatial parameters for gating of the bacterial large conductance mechanosensitive channel, MscL. J. Gen. Physiol. 113, 525­-540 (1999). 20. Sukharev, S. I., Martinac, B., Arshavsky, V. Y. & Kung, C. Two types of mechanosensitive channels in the Escherichia coli cell envelope: solubilization and functional reconstitution. Biophys. J. 65, 177­-183 (1993). 21. Ou, X. R., Blount, P., Hoffman, R. J. & Kung, C. One face of a transmembrane helix is crucial in mechanosensitive channel gating. Proc. Natl Acad. Sci. USA 95, 11471­-11475 (1998). 22. Maurer, J. A. & Dougherty, D. A. Generation and evaluation of a large mutational library from the Escherichia coli mechanosensitive channel of large conductance, MscL: implications for channel gating and evolutionary design. J. Biol. Chem. 278, 21076­-21082 (2003). 23. Cantor, R. S. Lateral pressures in cell membranes: a mechanism for modulation of protein function. J. Phys. Chem. 101, 1723­-1725 (1997). 24. Cantor, R. S. The influence of membrane lateral pressures on simple geometric models of protein conformational equilibria. Chem. Phys. Lipids 101, 45­-56 (1999). 25. Lindahl, E. & Edholm, O. Spatial and energetic-entropic decomposition of surface tension in lipid bilayers from molecular dynamics simulations. J. Chem. Phys. 113, 3882­-3893 (2000). 26. Gullingsrud, J. & Schulten, K. Gating of MscL studied by steered molecular dynamics. Biophys. J. 85, 2087­-2099 (2003). 27. Wiggins, P. & Phillips, R. Analytic models for mechanotransduction: gating a mechanosensitive channel. Proc. Natl Acad. Sci. USA 101, 4071­-4076 (2004). 28. Sheetz, M. P. & Singer, S. J. Biological membranes as bilayer couples. A molecular mechanism of drug-erythrocyte interactions. Proc. Natl Acad. Sci. USA 71, 4457­-4461 (1974). 29. Martinac, B., Adler, J. & Kung, C. Mechanosensitive ion channels of E. coli activated by amphipaths. Nature 348, 261­-263 (1990). 30. Lundbaek, J. A., Maer, A. M. & Andersen, O. S. Lipid bilayer electrostatic energy, curvature stress, and assembly of gramicidin channels. Biochemistry 36, 5695­-5701 (1997). 31. Cantor, R. S. The lateral pressure profile in membranes: a physical mechanism of general anesthesia. Toxicol. Lett. 100­-101, 451­-458 (1998). 32. Patel, A. J., Lazdunski, M. & Honore, E. Lipid and mechano-gated 2P domain K channels. Curr. Opin. Cell Biol. 13, 422­-428 (2001). 33. Patel, A. J. et al. A mammalian two pore domain mechano-gated S-like K channel. EMBO J. 17, 4283­-4290 (1998). 34. Maingret, F., Patel, A. J., Lesage, F., Lazdunski, M. & Honore, E. Lysophospholipids open the two-pore domain mechano-gated K channels TREK-1 and TRAAK. J. Biol. Chem. 275, 10128­-10133 (2000). 35. Patel, A. J. & Honore, E. Anesthetic-sensitive 2P domain K channels. Anesthesiology 95, 1013­-1021 (2001). 36. Chemin, J. et al. A phospholipid sensor controls mechanogating of the K channel TREK-1. EMBO J. 24, 44­-53 (2005). 37. Chemin, J. et al. Lysophosphatidic acid-operated K channels. J. Biol. Chem. 280, 4415­-4421 (2005). 38. Yang, X. C. & Sachs, F. Block of stretch-activated ion channels in Xenopus oocytes by gadolinium and calcium ions. Science 243, 1068­-1071 (1989). 39. Hamill, O. P. & McBride, D. W. Jr The pharmacology of mechanogated membrane ion channels. Pharmacol. Rev. 48, 231­-252 (1996). 40. Ermakov, Y. A., Averbakh, A. Z., Yusipovich, A. I. & Sukharev, S. Dipole potentials indicate restructuring of the membrane interface induced by gadolinium and beryllium ions. Biophys. J. 80, 1851­-1862 (2001). 41. Suchyna, T. M. et al. Identification of a peptide toxin from Grammostola spatulata spider venom that blocks cation-selective stretch-activated channels. J. Gen. Physiol. 115, 583­-598 (2000); erratum J. Gen. Physiol. 117, 371 (2001). 42. Guharay, G. & Sachs, F. Stretch-activated single ion channel currents in tissuecultured embryonic chick skeletal muscle. J. Physiol. (Lond.) 352, 685­-701 (1984). 43. Suchyna, T. M. et al. Bilayer-dependent inhibition of mechanosensitive channels by neuroactive peptide enantiomers. Nature 430, 235­-240 (2004). 44. Zhang, Y., Gao, F., Popov, V. L., Wen, J. W. & Hamill, O. P. Mechanically gated channel activity in cytoskeleton-deficient plasma membrane blebs and vesicles from Xenopus oocytes. J. Physiol. (Lond.) 523, 117­-130 (2000). 45. Maroto, R. et al. The role of TRPC1 in forming the mechanosensitive cation channel in frog oocytes. Nature Cell Biol. 7, 179­-185 (2005). 46. Morris, C. E. & Horn, R. Failure to elicit neuronal macroscopic mechanosensitive currents anticipated by single-channel studies. Science 251, 1246­-1249 (1991). 47. Zhang, Y. & Hamill, O. P. On the discrepancy between whole-cell and membrane patch mechanosensitivity in Xenopus oocytes. J. Physiol. (Lond.) 523, 101­-115 (2000). NATURE|Vol 436|4 August 2005 REVIEWS 653 2005 Nature Publishing Group 48. Ernstrom, G. G. & Chalfie, M. Genetics of sensory mechanotransduction. Annu. Rev. Genet. 36, 411­-453 (2002). 49. Goodman, M. B. & Schwarz, E. M. Transducing touch in Caenorhabditis elegans. Annu. Rev. Physiol. 65, 429­-452 (2003). 50. Bianchi, L. et al. The neurotoxic MEC-4(d) DEG/ENaC sodium channel conducts calcium: implications for nevrosis initiation. Nature Neurosci. 7, 1337­-1344 (2004). 51. Goodman, M. B. et al. MEC-2 regulates C. elegans DEG/ENaC channels needed for mechanosensation. Nature 415, 1039­-1042 (2002). 52. Chelur, D. S. et al. The mechanosensory protein MEC-6 is a subunit of the C. elegans touch-cell degenerin channel. Nature 420, 669­-673 (2002). 53. Emtage, L., Gu, G., Hartwieg, E. & Chalfie, M. Extracellular proteins organize the mechanosensory channel complex in C. elegans touch receptor neurons. Neuron 44, 795­-807 (2004). 54. Zhang, S. et al. MEC-2 is recruited to the putative mechanosensory complex in C. elegans touch receptor neurons through its stomatic-like domain. Curr. Biol. 14, 1888­-01896 (2004). 55. O'Hagan, R., Chalfie, M. & Goodman, M. B. The MEC-4 DEG/ENaC channel of Caenorhabditis elegans touch receptor neurons transduces mechanical signals. Nature Neurosci. 8, 43­-50 (2005). 56. Tsunoda, S. et al. A multivalent PDZ-domain protein assembles signalling complexes in a G-protein-coupled cascade. Nature 388, 243­-249 (1997). 57. Montell, C. TRP trapped in fly signalling web. Curr. Opin. Neurobiol. 8, 389­-397 (1998). 58. Montell, C., Birnbaumer, L. & Flockerzi, V. The TRP channels, a remarkably functional family. Cell 108, 595­-598 (2002). 59. Corey, D. P. New TRP channels in hearing and mechanosensation. Neuron 39, 585­-588 (2003). 60. Minke, B., Wu, C.-F. & Pak, W. L. Induction of photoreceptor voltage noise in the dark in Drosophila mutant. Nature 258, 84­-87 (1975). 61. Colbert, H. A., Smith, T. L. & Bargmann, C. I. OSM-9, a novel protein with structural similarity to channels, is required for olfaction, mechanosensation, and olfactory adaptation in Caenorhabditis elegans. J. Neurosci. 17, 8259­-8269 (1997). 62. Barr, M. M. & Sternberg, P. W. A polycystic kidney-disease gene homologue required for male mating behaviour in C. elegans. Nature 401, 386­-389 (1999). 63. Walker, R. G., Willingham, A. T. & Zuker, C. S. A Drosophila mechanosensory transduction channel. Science 287, 2229­-2234 (2000). 64. Tracey, W. D. Jr, Wilson, R. I., Laurent, G. & Benzer, S. painless, a Drosophila gene essential for nociception. Cell 113, 261­-273 (2003). 65. Di Palma, F. et al. Mutations in Mcoln3 associated with deafness and pigmentation defects in varitint-waddler (Va) mice. Proc. Natl Acad. Sci. USA 99, 14994­-14999 (2002). 66. Mochizuki, T. et al. PKD2, a gene for polycystic kidney disease that encodes an integral membrane protein. Science 272, 1339­-1342 (1996). 67. Suzuki, M., Mizuno, A., Kodaira, K. & Imai, M. Impaired pressure sensation in mice lacking TRPV4. J. Biol. Chem. 278, 22664­-22668 (2003). 68. Gao, X., Wu, L. & O'Neil, R. G. Temperature-modulated diversity of TRPV4 channel gating: activation by physical stresses and phorbol ester derivatives through protein kinase C-dependent and -independent pathways. J. Biol. Chem. 278, 27129­-27137 (2003). 69. Liedtke, W. et al. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 103, 525­-535 (2000). 70. Strotmann, R., Harteneck, C., Nunnenmacher, K., Schultz, G. & Plant, T. D. OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nature Cell Biol. 2, 695­-702 (2000). 71. Tobin, D. et al. Combinatorial expression of TRPV channel proteins defines their sensory functions and subcellular localization in C. elegans neurons. Neuron 35, 307­-318 (2002). 72. Kahn-Kirby, A. et al. Specific polyunsaturated fatty acids drive TRPV-dependent sensory signalling in vivo. Cell 119, 889­-900 (2004). 73. Liedtke, W., Tobin, D. M., Bargmann, C. I. & Friedman, J. M. Mammalian TRPV4 (VR-OAC) directs behavioural responses to osmotic and mechanical stimuli in Caenorhabditis elegans. Proc. Natl Acad. Sci. USA 100 (suppl. 2), 14531­-14536 (2003). 74. Watanabe, H. et al. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature 424, 434­-438 (2003). 75. Caterina, M. J. et al. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature 389, 816­-824 (1997). 76. Birder, L. A. et al. Altered urinary bladder function in mice lacking the vanilloid receptor TRPV1. Nature Neurosci. 5, 856­-860 (2002). 77. Vriens, J. et al. Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc. Natl Acad. Sci. USA 101, 396­-401 (2004). 78. Kim, J. et al. A TRPV family ion channel required for hearing in Drosophila. Nature 424, 81­-84 (2003). 79. Gong, S. et al. Two interdependent TRPV chanel subunits, Inactive and Nanchung, mediate hearing in Drosophila. J. Neurosci. 24, 9059­-9066 (2004). 80. Sidi, S., Friedrich, R. W. & Nicolson, T. NompC TRP channel required for vertebrate sensory hair cell mechanotransduction. Science 301, 96­-99 (2003). 81. Siemens, J. et al. Cadherin 23 is a component of the tip link in hair-cell stereocilia. Nature 428, 950­-955 (2004). 82. Sollner, C. et al. Mutations in cadherin 23 affect tip links in zebrafish sensory hair cells. Nature 428, 955­-959 (2004). 83. Corey, D. P. et al. TRPA1 is a candidate for the mechanosensitive transductin channel of vertebrate hair cells. Nature 432, 723­-730 (2004). 84. Howard, J. & Bechstedt, S. Hypothesis: a helix of ankyrin repeats of the NOMPC-TRP ion channel is the gating spring of mechanoreceptors. Curr. Biol. 14, R224­-R226 (2004). 85. Jiang, Y. et al. The open pore conformation of potassium channels. Nature 417, 523­-526 (2002). 86. Erler, I., Hirnet, D., Wissenbach, U., Flockerzi, V. & Niemeyer, B. A. Ca2 selective transient receptor potential V channel architecture and function require a specific ankyrin repeat. J. Biol. Chem. 279, 34456­-34463 (2004). 87. Hirono, M., Denis, C. S., Richardson, G. P. & Gillespie, P. G. Hair cells require phosphatidylinositol 4,5-bisphosphate for mechanical transduction and adaptation. Neuron 44, 309­-320 (2004). 88. Story, G. M. et al. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 112, 819­-829 (2003). 89. Jordt, S. E. et al. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature 427, 260­-265 (2004). 90. Bandell, M. et al. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron 41, 849­-857 (2004). 91. Palmer, C. P. et al. A TRP homolog in Saccharomyces cerevisiae forms an intracellular Ca2 -permeable channel in the yeast vacuolar membrane. Proc. Natl Acad. Sci. USA 98, 7801­-7805 (2001). 92. Zhou, X. L. et al. The transient receptor potential channel on the yeast vacuole is mechanosensitive. Proc. Natl Acad. Sci. USA 100, 7105­-7110 (2003). 93. Denis, V. & Cyert, M. S. Internal Ca(2 ) release in yeast is triggered by hypertonic shock and mediated by a TRP channel homologue. J. Cell Biol. 156, 29­-34 (2002). 94. Zhou, X.-L., Loukin, S. H., Coria, R., Kung, C. & Saimi, Y. Heterologously expressed fungal transient receptor potential channels retain mechanosensitivity in vitro and osmotic response in vivo. Eur. Biophys. J. 34, 413­-422 (2005). 95. Kung, C., Saimi, Y. & Martinac, B. Current Topics in Membranes and Transport 145­-153 (Academic, New York, 1990). 96. Apostle, H. G. Aristotle's On The Soul (De Anima) (Translation) 42­-43 (Peripatetic, Crinnell, Iowa, 1981). 97. Clapham, D. E. TRP channels as cellular sensors. Nature 426, 517­-524 (2003). 98. Voets, T. et al. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature 430, 748­-754 (2004). 99. Corey, D. P. & Sotomayor, M. Hearing: tightrope act. Nature 428, 901­-903 (2004). 100. Kung, C. in Evolution of the First Nervous Systems (ed. Anderson, P. A. V.) 203­-214 (Plenum, New York, 1990). Acknowledgements I thank A. Anishkin, M. Chalfie, R. Fettiplace, W. J. Haynes, S. Loukin, B. Martinac, Y. Saimi, A. O. W. Stretton and X.-L. Zhou for discussions and criticisms. My laboratory is supported by the Vilas Trust of the University of Wisconsin and by the NIH. Author Information Reprints and permissions information is available at npg.nature.com/reprintsandpermissions. The author declares no competing financial interests. Correspondence and requests for materials should be addressed to C.K. (ckung@wisc.edu). REVIEWS NATURE|Vol 436|4 August 2005 654 2005 Nature Publishing Group