
Keywords: calcium signaling, gastrointestinal motility, mechano-gated ion channels, mechanosensation, myogenic response
Abstract
Mechanosensation, the ability to properly sense mechanical stimuli and transduce them into physiologic responses, is an essential determinant of gastrointestinal (GI) function. Abnormalities in this process result in highly prevalent GI functional and motility disorders. In the GI tract, several cell types sense mechanical forces and transduce them into electrical signals, which elicit specific cellular responses. Some mechanosensitive cells like sensory neurons act as specialized mechanosensitive cells that detect forces and transduce signals into tissue-level physiological reactions. Nonspecialized mechanosensitive cells like smooth muscle cells (SMCs) adjust their function in response to forces. Mechanosensitive cells use various mechanoreceptors and mechanotransducers. Mechanoreceptors detect and convert force into electrical and biochemical signals, and mechanotransducers amplify and direct mechanoreceptor responses. Mechanoreceptors and mechanotransducers include ion channels, specialized cytoskeletal proteins, cell junction molecules, and G protein-coupled receptors. SMCs are particularly important due to their role as final effectors for motor function. Myogenic reflex—the ability of smooth muscle to contract in response to stretch rapidly—is a critical smooth muscle function. Such rapid mechanotransduction responses rely on mechano-gated and mechanosensitive ion channels, which alter their ion pores’ opening in response to force, allowing fast electrical and Ca2+ responses. Although GI SMCs express a variety of such ion channels, their identities remain unknown. Recent advancements in electrophysiological, genetic, in vivo imaging, and multi-omic technologies broaden our understanding of how SMC mechano-gated and mechanosensitive ion channels regulate GI functions. This review discusses GI SMC mechanosensitivity's current developments with a particular emphasis on mechano-gated and mechanosensitive ion channels.
INTRODUCTION
What Is Mechanotransduction?
Mechanotransduction is a process by which cells convert mechanical forces into electrical and chemical signals that result in specific cellular responses. Mechanotransduction is a two-step process. First, cells sense the mechanical input from their surroundings using mechanoreceptors. Second, cells convert the receptor signal into the appropriate cellular response using mechanotransducers. A cell may use a repertoire of mechanoreceptors: ion channels, transmembrane adhesion receptors, cell surface receptors, and sarcomeric proteins (1). Mechanotransduction is at the center of critical sensory processes, such as touch and hearing, and autonomic processes, such as blood pressure regulation, gastrointestinal (GI) motility, embryonic development, vascular tone, and muscle stretch (1–4). Mechanotransduction defects are associated with a range of genetic and acquired diseases, from muscular dystrophies, cardiomyopathies, and cancer, to GI functional and motility disorders (2, 5, 6).
Mechanotransduction in the GI Tract
Electromechanical organs like the GI tract, bladder, and heart generate tissue-specific mechanical forces sensed by mechanosensitive cells. In the GI tract, mechanotransduction is critical for normal function. Notable GI disorders that involve disrupted mechanotransduction include diverticulosis (7), ulcerative colitis (8), Crohn’s disease (8), chronic constipation (9), pseudo-obstruction, visceral hypersensitivity (10, 11), functional dyspepsia (12), irritable bowel syndrome, colon cancer (13), and even systemic conditions, like obesity (14). Since the GI tract is continually changing its volume and moving, all GI tract cells live in a mechanical world. We can classify the GI tract cells as specialized and nonspecialized mechanosensitive cells. Specialized mechanosensitive cells, like enterochromaffin cells, intrinsic primary afferent neurons, intestinofugal afferent neurons, and extrinsic sensory neurons, sense and convert mechanical forces into an appropriate physiological response. Meanwhile, nonspecialized mechanosensitive cells such as smooth muscle cells (SMCs), interstitial cells of Cajal (ICCs), and immune cells adjust their own function in response to mechanical stimuli (15–17).
SMC mechanotransduction is critical for normal GI tract development and subsequent function. During development, smooth muscle orients itself in response to mechanical forces, and this proper alignment of muscle layers is crucial for normal function (18). The role of SMC mechanotransduction extends and expands after development. It is critical for normal function since SMCs are the ultimate effectors—they generate resting tone, muscle contraction, and relaxation. As these functions require constant and rapid responses to mechanical stimuli, gut SMCs employ a wide array of membrane-associated molecules to sense and process the mechanical stimuli. These include surface receptors (ion channels, G protein-coupled receptors, and kinases), specialized intracellular cytoskeletal proteins, and extracellular cell-cell connections (1). Cells use these molecules to assimilate various mechanical stimuli like shear stress, stretch, pressure, torsion, and compression and convert them into short-term (i.e., changes in voltage and ion concentrations) and long-term (i.e., changes in gene expression) effects (5). The layers of the GI tract (mucosa, submucosa, muscularis, and serosa) have distinct mechanical properties and sense a range of passive and active mechanical forces, which ultimately result in efficient digestion, absorption, and propulsion of contents. Mechanical activity in the gut is wide ranging, including the peristaltic reflex, segmentation, propulsion, migratory motor activity, and structural rearrangements in response to function. In the muscle layer, the peristaltic, segmental, and ring contractions generate wall tension, whereas luminal content flow results in shear stress that drives the activation of various mechanosensitive components (19, 20).
Among the mechanically sensitive molecules, ion channels are indispensable for rapid mechanotransduction. Mechanically sensitive ion channels rapidly transduce force and quickly affect downstream mechanisms via electrical and calcium signaling. We may divide the mechanically sensitive ion channels into two categories. Mechano-gated ion channels are transmembrane proteins whose primary role is to respond to force (4, 21) and serve as classic mechanoreceptors. Mechanosensitive ion channels are gated by primary stimuli other than force, like chemical ligands and voltage, and force alters the channels' responses to their primary stimuli (22, 23). Mechano-gated ion channels sense forces, through either the membrane (force-from-lipid) (24) or structural proteins (force-from-tether) (25), by specialized mechanically sensitive gates that open (or close) in response to mechanical input (26). To be considered mechano-gated, an ion channel must meet specific criteria (4, 21): 1) It must be required for a response to mechanical stimuli and not for cellular development; 2) it must be expressed in a mechanosensitive cell, and the loss of the channel should abolish mechanical response; 3) alterations in the properties of the ion channel gene should alter the mechanical properties of the channel; and 4) it should retain the response to force when expressed in a heterologous system. The criteria for classifying the mechanosensitive ion channels do not currently exist. Eukaryotic mechano-gated and mechanosensitive channels' identities have been elusive, but recent advances in biophysics, bio-omics, and structural biology combined to substantially advance their identification (27–32). Indeed, the identification of mechano-gated ion channels is a critical step toward understanding the process of mechanotransduction. In this review, we highlight the importance of gut SMC mechanotransduction with a particular emphasis on SMC mechano-gated and mechanosensitive ion channels.
GI SMOOTH MUSCLE MECHANOSENSITIVITY
Myogenic Reflex as the Prototypical GI SMC Mechanotransduction
Coordinated contractility of GI SMCs is essential for normal GI motility. Synchronized interactions between myogenic, neurogenic, and hormonal mechanisms regulate GI smooth muscle tone and activity (33, 34). Neuronal and hormonal mechanisms adjust and synchronize motility patterns of the gut, whereas myogenic mechanisms may be fundamental processes since GI smooth muscle exhibits tone and contractility independent of neuronal input (33, 35).
In smooth muscle, the myogenic response, or reflex, refers to smooth muscle's ability to respond to mechanical forces independent from the nervous system. In a classic set of studies, Bayliss reported the contractile responses of arterial smooth muscle in response to increased transmural pressure (36). Although Edith Bulbring showed a myogenic response in GI smooth muscle (37), much of our mechanistic understanding of the myogenic response still derives from studies with vascular smooth muscle (38, 39). The smooth muscle syncytium consists of SMCs and interstitial cells (currently ICCs and PDGFR-α+ cells). ICCs exhibit spontaneous pacemaker activity in the stomach, small intestine, and colon, which is independent of hormonal and neuronal inputs. Pacemaker activity generates low-amplitude contractions that are further modulated by neuronal inputs (40). ICCs are mechanosensitive (16), and human ICCs express mechanosensitive ion channels such as NaV1.5 (124), but their role in GI mechanosensitivity remains poorly understood.
Ca2+ plays a crucial role in mechanotransduction through various mechanisms that involve changes in its cytosolic concentration. Ca2+ is essential for SMC function in the vascular system for maintaining arteriolar tone (41) and elicitation of myogenic constriction (38). The amplitudes and timing of Ca2+ signals are important signal transduction parameters. Rapid transient events or prolonged global responses are dependent on the intensity and timing of the stimulus (42). Mechano-gated ion channels are prime candidates for rapid Ca2+ entry in response to membrane force, and therefore, they are leading candidates to elicit the myogenic response. Various mechano-gated cation currents exist in vascular SMCs (43–45), carried by transient receptor potential (TRP) (39) and Piezo (46) channels. However, we know little about how GI SMCs contribute toward myogenic response, and even broadly, what molecular entities sense and mediate this response in gut SMCs.
Studies in toad stomach SMCs suggest that the opening of stretch-activated channels allows ions like Ca2+ to influx from the extracellular space, and subsequent membrane depolarization activates voltage-gated Ca2+ channels and Ca2+ release from internal stores that markedly raise intracellular Ca2+ (47). Small focal mechanical deformation of the plasma membrane of myocytes isolated from the rabbit distal colon's longitudinal muscle layer resulted in localized intracellular Ca2+ concentration transients that propagated throughout the cell (48). The mediators of mechanically induced Ca2+ influx may be L-type voltage-gated Ca2+ channels themselves (23), but since they are mechanosensitive, rather than mechano-gated, they may not be available to activate at SMC's resting potential. Further, in some experiments, Ca2+ channel blockers did not affect mechanically induced responses (38, 48), suggesting the involvement of other mechanically activated Ca2+ influx pathways and specifically mechano-gated cation channels. The cell membrane’s lipid composition and the interaction between the cytoskeleton and the extracellular membrane proteins may also influence the mechanosensitive ion channel activity (49–53). These studies demonstrate that SMCs exhibit diverse Ca2+ signaling patterns in response to mechanical stimuli, and mechano-gated ion channels take center stage in many rapid mechanosensing processes. In the subsequent sections, we discuss how different mechano-gated and mechanosensitive ion channels of gut SMCs might contribute toward gut myogenic response and mechanical sensitivity.
GI SMOOTH MUSCLE MECHANOSENSITIVE ION CHANNELS
In GI SMCs, the main candidates for rapid mechanotransduction required in myogenic response are mechanosensitive and mechano-gated ion channels (4, 52). Several types have been identified and characterized in GI SMCs, including but not limited to nonselective cation channels; voltage-gated Na+, Ca2+, and K+ channels; Ca2+-activated K+ channels; and two-pore K+ (K2P) channels (4, 38, 47) (Table 1).
Table 1.
Mechanosensitive and mechano-gated ion channels in GI SMCs
| Mechano Channel | Current | Type | Characterized | Location in GI Smooth Muscle | Reference |
|---|---|---|---|---|---|
| Cav1.2 | L-type Ca2+ | Mechanosensitive | Yes | Human jejunum circular SMCs, human stomach SMCs | (23, 54) |
| Nav1.5 | Na+ | Mechanosensitive | Yes | Human jejunal circular SMCs | (55) |
| Ca2+-activated large conductance K+ channel (BKCa) | K+ | Mechanosensitive | Yes | Rat colon SMCs | (56) |
| TREK-1 | K+ | Mechano-gated | Yes | Mouse small bowel and colon SMCs | (57–60) |
| TRPC4 | Nonselective cation | Mechanosensitive | Yes | Mouse stomach, small bowel, and colon SMCs | (61–64) |
| TRPC6 | Nonselective cation | Mechanosensitive | Yes | Mouse small bowel and colon SMCs | (61, 63,64) |
| TRPC7 | Nonselective cation | Mechanosensitive | No, RNA expression | Mouse small bowel and colon SMCs | (63,64) |
Ion Channels Whose Opening Depolarizes the Cell Membrane
Nonselective cation channels.
Membrane stretch activates nonselective cation channels in the cells from the cardiovascular system—rat atrial myocytes (65), guinea pig arterial SMCs (44), and porcine arterial SMCs (43). Similarly, pressurized patches from single SMCs from the toad stomach contained nonselective cation currents (47). Hypotonic cell swelling also increased carbachol-evoked inward current in guinea pig ileal SMCs and gastric myocytes through potentiation of nonselective cation channels (66, 67). Although the majority of nonselective mechano-gated cation channels are primarily permeable to Na+ and K+, they also conduct Ca2+ and Mg2+ (43, 66, 67). Indeed, single stretch-activated cationic channels in SMCs conduct quantities of Ca2+ adequate for stimulating the myogenic response (68). The commonly used nonselective mechano-gated ion channel blocker gadolinium (Gd3+) (69) eliminated the myogenic reflex and prevented action potential firing in stretched tissue (70). These studies suggest that nonselective cation channels may be important mediators of the myogenic reflex. Indeed, a variety of nonselective cation channels is found in GI SMCs (16).
Transient receptor potential channels.
The mammalian transient receptor potential (TRP) channel family is large. It comprises 28 members and is subdivided into five classes: TRPC, TRPV, TRPA, TRPP, and TRPM (71). TRP channels are cation-permeable ion channels gated by various stimuli, including chemicals like capsaicin (e.g., TRPV1) (72), temperature (e.g., TRPA1) (73), and force (e.g., TRPC6) (71). TRP channels are sensitive to various forms of mechanical force, including fluid shear stress and membrane stretch, and are potential candidates for mediating stretch-activated cation currents and pressure-induced depolarization (71). Among different subgroups, TRPC1/6, TRPM4, TRPV2/4, and TRPP1 (PKD2) have been associated with mechanotransduction in vascular SMCs (39, 71, 74, 75). Ca2+ influx through TRPC channels controls a variety of biological functions, including regulation of SMC proliferation, endothelin-evoked arterial contraction, neuronal differentiation, and cardiac hypertrophy (76). TRP ion channels are expressed in the GI tract, where they act as molecular sensors and transducers regulating various functions (61, 64, 77, 78). Deletion of TRPC4 and TRPC6 in GI SMCs results in impaired smooth muscle contraction and intestinal motility in vivo (63). Similarly, TRPC4 has been demonstrated as an essential component of the nonselective cation channels in murine stomach (62).
TRP channels are involved in a variety of mechanosensory processes. However, it remains unclear whether some TRP channels are the primary receptors for mechanical force (mechano-gated channels), or they are mechanosensitive or they are involved in downstream mechanotransduction. For example, we discuss the involvement of TRPC1 and TRPC6 channels in mechanosensing by a range of cells in the vascular tree. However, direct mechano-gating of TRPC1 and TRPC6 remains in question. Some studies suggest that they are mechano-gated, whereas others show that they are not directly mechano-gated when expressed in a heterologous system (79) or liposomes (80). It could be that cell-specific cofactors are required. TRPC6 expressed in Caenorhabditis elegans neurons restored mechanosensitivity in a touch-insensitive mutant but still required diacylglycerol for activation (80).
Piezo channels.
Piezo mechano-gated ion channels have recently been established as major force sensors in mammalian cells (28). The two isoforms, Piezo1 and Piezo2, are essential components of many mechanosensory processes during development (46, 81), homeostasis (82), blood pressure regulation (83), and cell and tissue restructuring in response to endogenous forces (46, 81). Piezo1 and Piezo2 mechano-gated ion channels are expressed throughout the length of the GI tract (28). Piezo2 is generally restricted to specialized mechanosensory cells—like epithelial Merkel cells and gut epithelial enteroendocrine cells (84), and neurons (85, 86). Piezo1 is more broadly expressed (2), and although it is expressed in vascular SMCs (46), its expression and roles in GI SMCs remain unknown. Piezo1 biophysical properties, including nonselective cationic permeability, fast activation, and slow-to-intermediate inactivation (28), make it a strong candidate to contribute to physiological functions such as the myogenic reflex. Piezo channels can be inhibited by the broad mechanosensitive channel blockers Gd3+ and ruthenium red, as well as the more specific Piezo channel inhibitor GsMTx-4 (15, 28, 46, 87). Indeed, guinea pig gastric SMCs express mechanically sensitive nonselective cation currents inhibited by Gd3+ and with kinetic and conductance properties like Piezo1 channels (88). Based on these findings, studies have begun exploring Piezo channels in the gut (89). However, the precise distribution of Piezo channels and their function in GI SMCs remains unexplored.
Cation-selective channels.
Voltage-gated ion channels regulate membrane potential and endow GI smooth muscle with electrical excitability, which is required for contractility. GI SMCs express voltage-gated Ca2+, Na+, and K+ channels, and some of these channels are mechanosensitive.
Voltage-gated Ca2+ channels.
Ca2+ influx into SMCs is essential for intestinal smooth muscle contractility. The main pathway for Ca2+ entry into SMCs is through the dihydropyridine-sensitive L-type Ca2+ channels. Calcium channels identified in human jejunum circular SMCs (CACNA1C-encoded CaV1.2) can be activated by an increase in positive pressure or external shear forces (23, 53) or by osmotic stress (54). In gastric SMCs, although f-actin cytoskeleton disruption diminished L-type currents in isotonic solutions, cytoskeletal disruption did not alter L-type currents in response to osmotic stress (54). Similarly, in rat myocytes cell swelling and application of negative or positive pressure increased the dihydropyridine-sensitive inward current. Interestingly, bilayer composition rather than intact cytoskeleton modifies L-type channel mechanosensitivity, leading to altered activity and Ca2+ entry in human jejunum circular layer myocytes (90).
The external muscle layer SMCs also express a low voltage-activated, Ca2+-permeable ionic conductance in mouse (91, 92) and human colon (93), as well as ICCs from dog and mouse colon, and mouse small intestine (94). Early studies in colonic myocytes characterized the currents as T-type (92), and later work identified T-type Ca2+ channel (Cacna1h-encoded CaV3.2) in mouse jejunum ICC and SMCs (91). CaV3.2 can be blocked by the gut antispasmodic otilonium bromide, and T-type blocker mibefradil can modulate human colonic smooth muscle contraction (95). Though T-type Ca2+ channels are involved in mechanotransduction (96), their mechanosensitivity and roles in GI SMCs remain unclear.
Voltage-gated Na+ channels.
SMCs are thought of as Ca2+-centric systems with minimal contribution from Na+ channels. However, this dogma has been successfully challenged, especially for human SMCs. SCN5A-encoded NaV1.5 is expressed in cardiomyocytes (97). Interestingly, there is species dependence, as mouse GI SMCs do not express NaV1.5 (55, 98, 99). Indeed, in human GI SMCs, the influx of Na+ through NaV1.5 quickly depolarizes the membrane potential, which was found to be important for the regulation of the electrical slow waves in GI smooth muscle (100). NaV1.5 is also mechanosensitive (22, 101, 102). Mechanical forces have several important effects on NaV1.5, including their gating, kinetics, and inactivation (22, 53). In human jejunum circular SMCs, NaV1.5 mechanosensitivity is diminished by f-actin disruption (53), and the mechanosensitive response of heterologously expressed NaV1.5 is regulated by the cytoskeletal proteins, syntrophin γ2 and telethonin (49, 50). The bilayer also has important effects on NaV1.5 mechanosensitivity. Lipid-permeable anesthetics, such as lidocaine and ranolazine, known to alter GI sensorimotor function, diminish NaV1.5 mechanosensitivity (9, 22, 103). Perhaps most importantly, human SCN5A mutations are associated with multiple clinical phenotypes, including conduction disorders in the heart and irritable bowel syndrome (IBS) in the GI tract (104, 105). IBS-associated SCN5A mutations frequently impact voltage gating (106) and/or mechanosensitivity (101, 102).
Ion Channels Whose Opening Hyperpolarizes the Cell Membrane
Several types of potassium channels exist in GI SMCs, including Ca2+-activated, voltage-gated, and two-pore K+ (K2P) channels. Variable K+ channel expression in SMCs contributes to the wide range of resting potentials and electrical patterns in GI smooth muscle (107). Except for inward-rectifying potassium channels, K+ channel opening results in K+ efflux, membrane hyperpolarization, and a consequent decrease in cell excitability (2). Therefore, although K+ channels are unlikely contributors to activation of the myogenic reflex, they may contribute as mechanical brakes, as recently shown in neurons (108) and vascular SMCs, the latter of which were activated during myogenic vasoconstriction to suppress response (109). They also contribute to active relaxation precipitated by organ filling, as in the proximal colon (58).
Ca2+-activated K+ channels.
GI SMCs express several Ca2+-activated K+ channels, which contribute to GI smooth muscle excitability (107). Mechanosensitive Ca2+-activated large conductance K+ channels (BKCa) produce large outward currents in GI SMCs, decreasing excitability (56). Compared to other GI segments, BKCa expression and current densities are highest in the colon (56). BKCa in rat colonic SMCs was not only sensitive to stretch, but BKCa blocker charybdotoxin attenuated the relaxation response of colonic smooth muscle strips to stretch (56).
Two-pore K+ channels.
Among the two-pore K+ (K2P) channel family, TREK-1 is specific to smooth muscle tissue in both mouse ileum and colon, whereas TREK-2 and TRAAK channels were found in enteric neurons but not in smooth muscle (58–60). In organ bath experiments, K2P channel activators induced the relaxation of ileum and colon tissues precontracted by KCl and carbachol and reduced the amplitude of spontaneous contractions (59). These channels contribute to membrane hyperpolarization and relaxation in the GI tract regions where relaxation is required upon mechanical stimulation, such as ascending colon that serves an important storage function (60). The K2P channel TREK-1 is mechano-gated (57). Interestingly, both dietary (51) and membrane lipids affect the mechanical sensitivity of several ion channels, and this is strongly established for TREK-1 (110). Studies have also demonstrated the presence of TREK-1 in rat bladder myocytes where the channel is activated by arachidonic acid and may have an important role in the regulation of bladder smooth muscle tone during urine storage (111).
Voltage-gated K+ channels.
Vascular SMCs express variety of voltage-gated K+ (KV) channel isoforms, which play important roles in regulating their contraction (3, 107), and several studies showed that KV channels are mechanosensitive (112, 113). In vascular SMCs, multiple KV channels contribute to resting membrane potential and exhibit negative feedback regulation of myogenic tone (3, 114, 115). GI SMCs also express several types of KV channels that contribute to the resting membrane potential in these cells (107, 116, 117). KV channel mechanosensitivity may serve as a “mechanical brake” to electrical excitation (108), but the roles of KV channel in SMC mechanosensitivity remain poorly understood.
To summarize, several types of mechano-gated and mechanosensitive ion channels are expressed in GI SMCs, and they regulate muscle contraction primarily through increasing the intracellular Ca2+ concentration. In many instances, these channels coordinate mechanotransduction. For example, TRPC1 and TRPC6 channels cooperate with TRPV4 in neurons to mediate mechanical hyperalgesia and primary afferent nociceptor sensitization (118). Similarly, TRPC3 and TRPC6 are co-expressed in sensory neuron cell lines, which are required for the normal function of cells involved in touch and hearing (119). Furthermore, activation of TRPV1 channels with capsaicin, in either dorsal root ganglion neurons or a heterologous expression system, inhibited Piezo1 and Piezo2 by depleting phosphatidylinositol 4,5-bisphosphate [PI(4,5)P]2 and its precursor PI(4)P from the plasma membrane through Ca2+-induced phospholipase Cδ (PLCδ) activation (120). ENaC, TRPM4, and TRPC6 also play important roles in the pressure-induced myogenic response, and ENaC and TRPM4 interact in rat posterior cerebral arteries (121). Piezo1 and Piezo2 channels are rarely co-expressed in the same cell, but when they are as in baroreceptors (122) and chondrocytes (123), they collaborate to produce a physiologic response.
CONCLUSIONS AND FUTURE DIRECTIONS
Coordinated myogenic and neurogenic mechanisms regulate smooth muscle function and hence gut motility. Smooth muscle contraction in response to mechanical force occurs after blocking neuronal activity, a phenomenon called myogenic reflex. However, the mechanistic understanding of the myogenic reflex and other mechanosensing behavior of SMCs is derived mostly from the vascular system, where these concepts are deeply explored, whereas these mechanisms in GI SMCs remain mostly undetermined.
GI SMCs express a wide range of mechanosensitive ion channels. Like vascular SMCs, mechanosensitive ion channels present in GI SMCs may contribute to the myogenic response, but their identities and roles remain unclear. There are important, but not insurmountable, barriers to the progress in this area. Perhaps the biggest hurdle is the identification of mechano-gated channels. As we described above, an ion channel needs to fulfill certain criteria to be considered mechano-gated (4, 21) or mechanosensitive. Several ion channels initially considered as candidate mechano-gated channels failed to fulfill some of these criteria. A classic example is the TRP channel family. Many TRP channels were initially considered as mechano-gated (124, 125). However, recent studies demonstrated that many of these TRP channels are not mechano-gated (79). Instead, they either are involved in downstream mechanotransduction processes or require additional cellular components like the cytoskeleton to function as mechanoreceptors (124, 126–128). To define the identities and roles of mechano-gated channels in mechanosensitive cells like GI SMCs, we must clearly define the mechanically modulated channels as mechano-gated or mechanosensitive, and we need to pay close attention to these channels' subcellular localization and mechanical nano-environment. These data will be required to determine how GI SMCs integrate different mechanical inputs; whether force is distributed between the lipid bilayer, cytoskeletal proteins, and cell surface mechano-gated channels; whether different gating mechanisms are involved for a specific mechanical input; and how different types of mechanosensitive ion channels (depolarizing and hyperpolarizing) are regulated. Recent developments in function-transcription approaches, single-cell RNA sequencing, and spatial transcriptomics technologies are giving new opportunities in understanding the biochemical processes at the cellular level. Single-cell RNA sequencing of vascular SMCs has yielded knowledge about their heterogeneity (129), ability to phenotypically modulate in response to various stimuli (130, 131), and developmental pathways (132, 133). Further, single-cell RNA sequencing following whole cell electrophysiology (Patch-seq) shed light on functional diversity in excitable cell types (134–137). Although single-cell RNA sequence analysis gives valuable information regarding cellular heterogeneity, spatial transcriptomics renders spatially defined characteristics of these cells. Combination of single-cell sequence and spatial analysis simultaneously provides cellular heterogeneity and their spatial distribution and association with each other in a tissue (138). Applying these techniques in GI SMCs will help to understand distinct smooth muscle subtypes in different regions of the GI tract, their functional diversity, and how they integrate with other cell types to exhibit physiologic functions, like the classic myogenic reflex riddle.
The progress in our understanding of the GI SMC myogenic response is further hampered by species, sex, age, and regional differences in their mechanotransduction mechanisms. GI SMCs are not a homogenous population, and they have varying morphological (139), electrical, and mechanical properties (40). Further, myogenic properties vary in different regions of the GI tract and between circular and longitudinal muscle layers, and these systems are likely optimized to perform the requisite functions. Apparent differences in myogenic responses also exist between species and with age. Notable example abounds. Na+ currents carried by NaV1.5 were identified in human and dog SMCs, but not in mouse pig or guinea pig SMCs (99), and myogenic peristaltic contractions can be stimulated by muscarinic agonists in isolated rabbit and rat colon, but not in guinea pig or mouse colon (140). Age is a critical factor as well. For example, in colonic smooth muscle tissue of baboons, carbachol-induced contractile responses were reduced with age (141), and changes in gut anatomy and function with age are deeply established (142).
Smooth muscles are capable of autoregulation and actively contract in response to stretch. Ion channels modulated by mechanical forces likely contribute to the initiation and transduction of the myogenic response. Studies in the vascular SMCs and other cell types have identified several potential ion channels involved in mechanotransduction. However, knowledge regarding GI SMCs mechanotransduction is still limited. Further research on mechano-gated and mechanosensitive ion channel complexes of GI SMCs, along with their molecular mechanisms, distribution, and coregulation/colocalization with other cellular components in the GI tract, will advance our understanding of smooth muscle tissue and help to develop novel treatment strategies for GI motility disorders.
GRANTS
The authors are supported by National Institutes of Health Grants DK106456, DK123549, AT010875, and DK052766.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
A.B. conceived and designed research; V.J. and P.R.S. drafted manuscript; V.J., P.R.S., G.F., and A.B. edited and revised manuscript; V.J., P.R.S., G.F., and A.B. approved final version of manuscript.
ACKNOWLEDGMENTS
We thank Dr. Szurszewski and Arnaldo Mercado-Perez for constructive feedback and Kristy Zodrow for administrative assistance.
REFERENCES
- 1.Eyckmans J, Boudou T, Yu X, Chen CS. A hitchhiker’s guide to mechanobiology. Dev Cell 21: 35–47, 2011. doi: 10.1016/j.devcel.2011.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Douguet D, Honore E. Mammalian mechanoelectrical transduction: structure and function of force-gated ion channels. Cell 179: 340–354, 2019. doi: 10.1016/j.cell.2019.08.049. [DOI] [PubMed] [Google Scholar]
- 3.Jackson WF. KV channels and the regulation of vascular smooth muscle tone. Microcirculation 25: e12421, 2018. doi: 10.1111/micc.12421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Ranade SS, Syeda R, Patapoutian A. Mechanically activated ion channels. Neuron 87: 1162–1179, 2015. [Erratum in Neuron 88: 433, 2015]. doi: 10.1016/j.neuron.2015.08.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Delmas P, Coste B. Mechano-gated ion channels in sensory systems. Cell 155: 278–284, 2013. doi: 10.1016/j.cell.2013.09.026. [DOI] [PubMed] [Google Scholar]
- 6.Jaalouk DE, Lammerding J. Mechanotransduction gone awry. Nat Rev Mol Cell Biol 10: 63–73, 2009. doi: 10.1038/nrm2597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Patel B, Guo X, Noblet J, Chambers S, Gregersen H, Kassab GS. Computational analysis of mechanical stress in colonic diverticulosis. Sci Rep 10: 6014, 2020. doi: 10.1038/s41598-020-63049-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Moriggi M, Pastorelli L, Torretta E, Tontini GE, Capitanio D, Bogetto SF, Vecchi M, Gelfi C. Contribution of extracellular matrix and signal mechanotransduction to epithelial cell damage in inflammatory bowel disease patients: a proteomic study. Proteomics 17: 1700164, 2017. doi: 10.1002/pmic.201700164. [DOI] [PubMed] [Google Scholar]
- 9.Neshatian L, Strege PR, Rhee PL, Kraichely RE, Mazzone A, Bernard CE, Cima RR, Larson DW, Dozois EJ, Kline CF, Mohler PJ, Beyder A, Farrugia G. Ranolazine inhibits voltage-gated mechanosensitive sodium channels in human colon circular smooth muscle cells. Am J Physiol Gastrointest Liver Physiol 309: G506–G512, 2015. doi: 10.1152/ajpgi.00051.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Balemans D, Boeckxstaens GE, Talavera K, Wouters MM. Transient receptor potential ion channel function in sensory transduction and cellular signaling cascades underlying visceral hypersensitivity. Am J Physiol Gastrointest Liver Physiol 312: G635–G648, 2017. doi: 10.1152/ajpgi.00401.2016. [DOI] [PubMed] [Google Scholar]
- 11.Feng B, Guo T. Visceral pain from colon and rectum: the mechanotransduction and biomechanics. J Neural Transm (Vienna) 127: 415–429, 2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Page AJ, Li H. Meal-sensing signaling pathways in functional dyspepsia. Front Syst Neurosci 12, 2018. doi: 10.3389/fnsys.2018.0001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Fernandez-Sanchez ME, Barbier S, Whitehead J, Bealle G, Michel A, Latorre-Ossa H, Rey C, Fouassier L, Claperon A, Brulle L, Girard E, Servant N, Rio-Frio T, Marie H, Lesieur S, Housset C, Gennisson JL, Tanter M, Menager C, Fre S, Robine S, Farge E. Mechanical induction of the tumorigenic beta-catenin pathway by tumour growth pressure. Nature 523: 92–95, 2015. doi: 10.1038/nature14329. [DOI] [PubMed] [Google Scholar]
- 14.Acosta A, Camilleri M, Shin A, Vazquez-Roque MI, Iturrino J, Burton D, O'Neill J, Eckert D, Zinsmeister AR. Quantitative gastrointestinal and psychological traits associated with obesity and response to weight-loss therapy. Gastroenterology 148: 537–546, 2015. doi: 10.1053/j.gastro.2014.11.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Alcaino C, Farrugia G, Beyder A. Mechanosensitive piezo channels in the gastrointestinal tract. Curr Top Membr 79: 219–244, 2017. doi: 10.1016/bs.ctm.2016.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Kraichely RE, Farrugia G. Mechanosensitive ion channels in interstitial cells of Cajal and smooth muscle of the gastrointestinal tract. Neurogastroenterol Motil 19: 245–252, 2007. doi: 10.1111/j.1365-2982.2006.00880.x. [DOI] [PubMed] [Google Scholar]
- 17.Mazet B. Gastrointestinal motility and its enteric actors in mechanosensitivity: past and present. Pflugers Arch 467: 191–200, 2015. doi: 10.1007/s00424-014-1635-7. [DOI] [PubMed] [Google Scholar]
- 18.Huycke TR, Miller BM, Gill HK, Nerurkar NL, Sprinzak D, Mahadevan L, Tabin CJ. Genetic and mechanical regulation of intestinal smooth muscle development. Cell 179: 90–105, 2019. doi: 10.1016/j.cell.2019.08.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Beyder A. In pursuit of the epithelial mechanosensitivity mechanisms. Front Endocrinol (Lausanne) 9: 804, 2018. doi: 10.3389/fendo.2018.00804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Gayer CP, Basson MD. The effects of mechanical forces on intestinal physiology and pathology. Cell Signal 21: 1237–1244, 2009. doi: 10.1016/j.cellsig.2009.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Arnadottir J, Chalfie M. Eukaryotic mechanosensitive channels. Annu Rev Biophys 39: 111–137, 2010. doi: 10.1146/annurev.biophys.37.032807.125836. [DOI] [PubMed] [Google Scholar]
- 22.Beyder A, Rae JL, Bernard C, Strege PR, Sachs F, Farrugia G. Mechanosensitivity of Nav1.5, a voltage-sensitive sodium channel. J Physiol 588: 4969–4985, 2010. doi: 10.1113/jphysiol.2010.199034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Farrugia G, Holm AN, Rich A, Sarr MG, Szurszewski JH, Rae JL. A mechanosensitive calcium channel in human intestinal smooth muscle cells. Gastroenterology 117: 900–905, 1999. doi: 10.1016/s0016-5085(99)70349-5. [DOI] [PubMed] [Google Scholar]
- 24.Anishkin A, Kung C. Stiffened lipid platforms at molecular force foci. Proc Natl Acad Sci USA 110: 4886–4892, 2013. doi: 10.1073/pnas.1302018110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Zhang W, Cheng LE, Kittelmann M, Li J, Petkovic M, Cheng T, Jin P, Guo Z, Gopfert MC, Jan LY, Jan YN. Ankyrin repeats convey force to gate the NOMPC mechanotransduction channel. Cell 162: 1391–1403, 2015. doi: 10.1016/j.cell.2015.08.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Kefauver JM, Ward AB, Patapoutian A. Discoveries in structure and physiology of mechanically activated ion channels. Nature 587: 567–576, 2020. doi: 10.1038/s41586-020-2933-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Beaulieu-Laroche L, Christin M, Donoghue A, Agosti F, Yousefpour N, Petitjean H, Davidova A, Stanton C, Khan U, Dietz C, Faure E, Fatima T, MacPherson A, Mouchbahani-Constance S, Bisson DG, Haglund L, Ouellet JA, Stone LS, Samson J, Smith MJ, Ask K, Ribeiro-da-Silva A, Blunck R, Poole K, Bourinet E, Sharif-Naeini R. TACAN is an ion channel involved in sensing mechanical pain. Cell 180: 956–967, 2020. doi: 10.1016/j.cell.2020.01.033. [DOI] [PubMed] [Google Scholar]
- 28.Coste B, Mathur J, Schmidt M, Earley TJ, Ranade S, Petrus MJ, Dubin AE, Patapoutian A. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 330: 55–60, 2010. doi: 10.1126/science.1193270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Hong GS, Lee B, Wee J, Chun H, Kim H, Jung J, Cha JY, Riew TR, Kim GH, Kim IB, Oh U. Tentonin 3/TMEM150c confers distinct mechanosensitive currents in dorsal-root ganglion neurons with proprioceptive function. Neuron 91: 107–118, 2016. doi: 10.1016/j.neuron.2016.05.029. [DOI] [PubMed] [Google Scholar]
- 30.Maingret F, Patel AJ, 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. doi: 10.1074/jbc.275.14.10128. [DOI] [PubMed] [Google Scholar]
- 31.Murthy SE, Dubin AE, Whitwam T, Jojoa-Cruz S, Cahalan SM, Mousavi SAR, Ward AB, Patapoutian A. OSCA/TMEM63 are an evolutionarily conserved family of mechanically activated ion channels. eLife 7: e41844, 2018. doi: 10.7554/eLife.41844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Patkunarajah A, Stear JH, Moroni M, Schroeter L, Blaszkiewicz J, Tearle JL, Cox CD, Furst C, Sanchez-Carranza O, Ocana FM, Fleischer R, Eravci M, Weise C, Martinac B, Biro M, Lewin GR, Poole K. TMEM87a/Elkin1, a component of a novel mechanoelectrical transduction pathway, modulates melanoma adhesion and migration. eLife 9: e53308, 2020. doi: 10.7554/eLife.53308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Bortoff A. Myogenic control of intestinal motility. Physiol Rev 56: 418–434, 1976. doi: 10.1152/physrev.1976.56.2.418. [DOI] [PubMed] [Google Scholar]
- 34.Spencer NJ, Dinning PG, Brookes SJ, Costa M. Insights into the mechanisms underlying colonic motor patterns. J Physiol 594: 4099–4116, 2016. doi: 10.1113/JP271919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Duthie HL. Electrical activity of gastrointestinal smooth muscle. Gut 15: 669–681, 1974. doi: 10.1136/gut.15.8.669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Bayliss WM. On the local reactions of the arterial wall to changes of internal pressure. J Physiol 28: 220–231, 1902. doi: 10.1113/jphysiol.1902.sp000911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Bulbring E. Correlation between membrane potential, spike discharge and tension in smooth muscle. J Physiol 128: 200–221, 1955. doi: 10.1113/jphysiol.1955.sp005299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Davis MJ, Meininger GA, Zawieja DC. Stretch-induced increases in intracellular calcium of isolated vascular smooth muscle cells. Am J Physiol Heart Circ Physiol 263: H1292–H1299, 1992. doi: 10.1152/ajpheart.1992.263.4.H1292. [DOI] [PubMed] [Google Scholar]
- 39.Hill-Eubanks DC, Gonzales AL, Sonkusare SK, Nelson MT. Vascular TRP channels: performing under pressure and going with the flow. Physiology (Bethesda) 29: 343–360, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Sanders KM, Koh SD, Ro S, Ward SM. Regulation of gastrointestinal motility–insights from smooth muscle biology. Nat Rev Gastroenterol Hepatol 9: 633–645, 2012. doi: 10.1038/nrgastro.2012.168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Uchida E, Bohr DF. Myogenic tone in isolated perfused resistance vessels from rats. Am J Physiol 216: 1343–1350, 1969. doi: 10.1152/ajplegacy.1969.216.6.1343. [DOI] [PubMed] [Google Scholar]
- 42.Zhu MX, Ma J, Parrington J, Calcraft PJ, Galione A, Evans AM. Calcium signaling via two-pore channels: local or global, that is the question. Am J Physiol Cell Physiol 298: C430–C441, 2010. doi: 10.1152/ajpcell.00475.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Davis MJ, Donovitz JA, Hood JD. Stretch-activated single-channel and whole cell currents in vascular smooth muscle cells. Am J Physiol Cell Physiol 262: C1083–C1088, 1992. doi: 10.1152/ajpcell.1992.262.4.C1083. [DOI] [PubMed] [Google Scholar]
- 44.Setoguchi M, Ohya Y, Abe I, Fujishima M. Stretch-activated whole-cell currents in smooth muscle cells from mesenteric resistance artery of guinea-pig. J Physiol 501: 343–353, 1997. doi: 10.1111/j.1469-7793.1997.343bn.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Wu X, Davis MJ. Characterization of stretch-activated cation current in coronary smooth muscle cells. Am J Physiol Heart Circ Physiol 280: H1751–H1761, 2001. doi: 10.1152/ajpheart.2001.280.4.H1751. [DOI] [PubMed] [Google Scholar]
- 46.Retailleau K, Duprat F, Arhatte M, Ranade SS, Peyronnet R, Martins JR, Jodar M, Moro C, Offermanns S, Feng Y, Demolombe S, Patel A, Honore E. Piezo1 in smooth muscle cells is involved in hypertension-dependent arterial remodeling. Cell Rep 13: 1161–1171, 2015. doi: 10.1016/j.celrep.2015.09.072. [DOI] [PubMed] [Google Scholar]
- 47.Kirber MT, Guerrero-Hernandez A, Bowman DS, Fogarty KE, Tuft RA, Singer JJ, Fay FS. Multiple pathways responsible for the stretch-induced increase in Ca2+ concentration in toad stomach smooth muscle cells. J Physiol 524: 3–17, 2000. doi: 10.1111/j.1469-7793.2000.t01-4-00003.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Young SH, Ennes HS, McRoberts JA, Chaban VV, Dea SK, Mayer EA. Calcium waves in colonic myocytes produced by mechanical and receptor-mediated stimulation. Am J Physiol Gsatrointest Liver Physiol 276: G1204–G1212, 1999. doi: 10.1152/ajpgi.1999.276.5.G1204. [DOI] [PubMed] [Google Scholar]
- 49.Mazzone A, Strege PR, Tester DJ, Bernard CE, Faulkner G, De Giorgio R, Makielski JC, Stanghellini V, Gibbons SJ, Ackerman MJ, Farrugia G. A mutation in telethonin alters Nav1.5 function. J Biol Chem 283: 16537–16544, 2008. [Erratum in J Biol Chem 283: 22336, 2008]. doi: 10.1074/jbc.M801744200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Ou Y, Strege P, Miller SM, Makielski J, Ackerman M, Gibbons SJ, Farrugia G. Syntrophin gamma 2 regulates SCN5A gating by a PDZ domain-mediated interaction. J Biol Chem 278: 1915–1923, 2003. doi: 10.1074/jbc.M209938200. [DOI] [PubMed] [Google Scholar]
- 51.Romero LO, Massey AE, Mata-Daboin AD, Sierra-Valdez FJ, Chauhan SC, Cordero-Morales JF, Vasquez V. Dietary fatty acids fine-tune Piezo1 mechanical response. Nat Commun 10: 1200, 2019. doi: 10.1038/s41467-019-09055-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Sachs F, Morris CE. Mechanosensitive ion channels in nonspecialized cells. Rev Physiol Biochem Pharmacol 132: 1–77, 1998. doi: 10.1007/BFb0004985. [DOI] [PubMed] [Google Scholar]
- 53.Strege PR, Holm AN, Rich A, Miller SM, Ou Y, Sarr MG, Farrugia G. Cytoskeletal modulation of sodium current in human jejunal circular smooth muscle cells. Am J Physiol Cell Physiol 284: C60–C66, 2003. doi: 10.1152/ajpcell.00532.2001. [DOI] [PubMed] [Google Scholar]
- 54.Kim JH, Rhee PL, Kang TM. Actin cytoskeletons regulate the stretch-induced increase of Ca2+ current in human gastric myocytes. Biochem Biophys Res Commun 352: 503–508, 2007. doi: 10.1016/j.bbrc.2006.11.051. [DOI] [PubMed] [Google Scholar]
- 55.Ou Y, Gibbons SJ, Miller SM, Strege PR, Rich A, Distad MA, Ackerman MJ, Rae JL, Szurszewski JH, Farrugia G. SCN5A is expressed in human jejunal circular smooth muscle cells. Neurogastroenterol Motil 14: 477–486, 2002. doi: 10.1046/j.1365-2982.2002.00348.x. [DOI] [PubMed] [Google Scholar]
- 56.Ren J, Xin F, Liu P, Zhao HY, Zhang ST, Han P, Huang HX, Wang W. Role of BKCa in stretch-induced relaxation of colonic smooth muscle. Biomed Res Int 2016: 9497041, 2016. doi: 10.1155/2016/9497041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Berrier C, Pozza A, de Lacroix de Lavalette A, Chardonnet S, Mesneau A, Jaxel C, Le Maire M, Ghazi A. The purified mechanosensitive channel TREK-1 is directly sensitive to membrane tension. J Biol Chem 288: 27307–27314, 2013. doi: 10.1074/jbc.M113.478321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Koh SD, Sanders KM. Stretch-dependent potassium channels in murine colonic smooth muscle cells. J Physiol 533: 155–163, 2001. doi: 10.1111/j.1469-7793.2001.0155b.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Ma R, Seifi M, Papanikolaou M, Brown JF, Swinny JD, Lewis A. TREK-1 channel expression in smooth muscle as a target for regulating murine intestinal contractility: therapeutic implications for motility disorders. Front Physiol 9: 157, 2018. doi: 10.3389/fphys.2018.00157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Sanders KM, Koh SD. Two-pore-domain potassium channels in smooth muscles: new components of myogenic regulation. J Physiol 570: 37–43, 2006. doi: 10.1113/jphysiol.2005.098897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Gonzalez-Cobos JC, Trebak M. TRPC channels in smooth muscle cells. Front Biosci (Landmark Ed) 15: 1023–1039, 2010. doi: 10.2741/3660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Lee KP, Jun JY, Chang IY, Suh SH, So I, Kim KW. TRPC4 is an essential component of the nonselective cation channel activated by muscarinic stimulation in mouse visceral smooth muscle cells. Mol Cells 20: 435–441, 2005. [PubMed] [Google Scholar]
- 63.Tsvilovskyy VV, Zholos AV, Aberle T, Philipp SE, Dietrich A, Zhu MX, Birnbaumer L, Freichel M, Flockerzi V. Deletion of TRPC4 and TRPC6 in mice impairs smooth muscle contraction and intestinal motility in vivo. Gastroenterology 137: 1415–1424, 2009. doi: 10.1053/j.gastro.2009.06.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Walker RL, Hume JR, Horowitz B. Differential expression and alternative splicing of TRP channel genes in smooth muscles. Am J Physiol Cell Physiol 280: C1184–C1192, 2001. doi: 10.1152/ajpcell.2001.280.5.C1184. [DOI] [PubMed] [Google Scholar]
- 65.Zhang YH, Youm JB, Sung HK, Lee SH, Ryu SY, Ho WK, Earm YE. Stretch-activated and background non-selective cation channels in rat atrial myocytes. J Physiol 523: 607–619, 2000. doi: 10.1111/j.1469-7793.2000.00607.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Li L, Jin NG, Piao L, Hong MY, Jin ZY, Li Y, Xu WX. Hyposmotic membrane stretch potentiated muscarinic receptor agonist-induced depolarization of membrane potential in guinea-pig gastric myocytes. World J Gastroenterol 8: 724–727, 2002. doi: 10.3748/wjg.v8.i4.724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Waniishi Y, Inoue R, Ito Y. Preferential potentiation by hypotonic cell swelling of muscarinic cation current in guinea pig ileum. Am J Physiol 272: C240–C253, 1997. [DOI] [PubMed] [Google Scholar]
- 68.Zou H, Lifshitz LM, Tuft RA, Fogarty KE, Singer JJ. Visualization of Ca2+ entry through single stretch-activated cation channels. Proc Natl Acad Sci USA 99: 6404–6409, 2002. doi: 10.1073/pnas.092654999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Yang XC, Sachs F. Block of stretch-activated ion channels in Xenopus oocytes by gadolinium and calcium ions. Science 243: 1068–1071, 1989. doi: 10.1126/science.2466333. [DOI] [PubMed] [Google Scholar]
- 70.Kunze WA, Clerc N, Bertrand PP, Furness JB. Contractile activity in intestinal muscle evokes action potential discharge in guinea-pig myenteric neurons. J Physiol 517: 547–561, 1999. doi: 10.1111/j.1469-7793.1999.0547t.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Earley S, Brayden JE. Transient receptor potential channels in the vasculature. Physiol Rev 95: 645–690, 2015. doi: 10.1152/physrev.00026.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Song S, Ayon RJ, Yamamura A, Yamamura H, Dash S, Babicheva A, Tang H, Sun X, Cordery AG, Khalpey Z, Black SM, Desai AA, Rischard F, McDermott KM, Garcia JG, Makino A, Yuan JX. Capsaicin-induced Ca2+ signaling is enhanced via upregulated TRPV1 channels in pulmonary artery smooth muscle cells from patients with idiopathic PAH. Am J Physiol Lung Cell Mol Physiol 312: L309–L325, 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Aubdool AA, Graepel R, Kodji X, Alawi KM, Bodkin JV, Srivastava S, Gentry C, Heads R, Grant AD, Fernandes ES, Bevan S, Brain SD. TRPA1 is essential for the vascular response to environmental cold exposure. Nat Commun 5: 5732, 2014. doi: 10.1038/ncomms6732. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Nauli SM, Zhou J. Polycystins and mechanosensation in renal and nodal cilia. Bioessays 26: 844–856, 2004. [DOI] [PubMed] [Google Scholar]
- 75.Soni H, Peixoto-Neves D, Matthews AT, Adebiyi A. TRPV4 channels contribute to renal myogenic autoregulation in neonatal pigs. Am J Physiol Renal Physiol 313: F1136–F1148, 2017. doi: 10.1152/ajprenal.00300.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Rychkov G, Barritt GJ. TRPC1 Ca(2+)-permeable channels in animal cells. In: Transient Receptor Potential (TRP) Channels. Handbook of Experimental Pharmacology, edited by Flockerzi V, Nilius B. vol 179. Berlin: Springer. 2007, p. 23–52. doi: 10.1007/978-3-540-34891-7_2. [DOI] [PubMed] [Google Scholar]
- 77.Blackshaw LA, Brierley SM, Hughes PA. TRP channels: new targets for visceral pain. Gut 59: 126–135, 2010. [DOI] [PubMed] [Google Scholar]
- 78.Boesmans W, Owsianik G, Tack J, Voets T, Vanden Berghe P. TRP channels in neurogastroenterology: opportunities for therapeutic intervention. Br J Pharmacol 162: 18–37, 2011. doi: 10.1111/j.1476-5381.2010.01009.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Gottlieb P, Folgering J, Maroto R, Raso A, Wood TG, Kurosky A, Bowman C, Bichet D, Patel A, Sachs F, Martinac B, Hamill OP, Honore E. Revisiting TRPC1 and TRPC6 mechanosensitivity. Pflugers Arch 455: 1097–1103, 2008. doi: 10.1007/s00424-007-0359-3. [DOI] [PubMed] [Google Scholar]
- 80.Nikolaev YA, Cox CD, Ridone P, Rohde PR, Cordero-Morales JF, Vasquez V, Laver DR, Martinac B. Mammalian TRP ion channels are insensitive to membrane stretch. J Cell Sci 132: jcs238360, 2019. doi: 10.1242/jcs.238360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Ranade SS, Qiu Z, Woo SH, Hur SS, Murthy SE, Cahalan SM, Xu J, Mathur J, Bandell M, Coste B, Li YS, Chien S, Patapoutian A. Piezo1, a mechanically activated ion channel, is required for vascular development in mice. Proc Natl Acad Sci USA 111: 10347–10352, 2014. doi: 10.1073/pnas.1409233111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Faucherre A, Kissa K, Nargeot J, Mangoni ME, Jopling C. Piezo1 plays a role in erythrocyte volume homeostasis. Haematologica 99: 70–75, 2014. doi: 10.3324/haematol.2013.086090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Beech DJ, Kalli AC. Force sensing by piezo channels in cardiovascular health and disease. Arterioscler Thromb Vasc Biol 39: 2228–2239, 2019. doi: 10.1161/ATVBAHA.119.313348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Alcaino C, Knutson K, Treichel AJ, Yildiz G, Strege PR, Linden DR, Li JH, Leiter AB, Szurszewski JH, Farrugia G, Beyder A. A population of gut epithelial enterochromaffin cells is mechanosensitive and requires Piezo2 to convert force into serotonin release. Proc Natl Acad Sci USA 115: E7632–E7641, 2018. doi: 10.1073/pnas.1804938115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Nonomura K, Woo SH, Chang RB, Gillich A, Qiu Z, Francisco AG, Ranade SS, Liberles SD, Patapoutian A. Piezo2 senses airway stretch and mediates lung inflation-induced apnoea. Nature 541: 176–181, 2017. doi: 10.1038/nature20793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Ranade SS, Woo SH, Dubin AE, Moshourab RA, Wetzel C, Petrus M, Mathur J, Begay V, Coste B, Mainquist J, Wilson AJ, Francisco AG, Reddy K, Qiu Z, Wood JN, Lewin GR, Patapoutian A. Piezo2 is the major transducer of mechanical forces for touch sensation in mice. Nature 516: 121–125, 2014. doi: 10.1038/nature13980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Bae C, Sachs F, Gottlieb PA. The mechanosensitive ion channel Piezo1 is inhibited by the peptide GsMTx4. Biochemistry 50: 6295–6300, 2011. doi: 10.1021/bi200770q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Yamamoto Y, Suzuki H. Two types of stretch-activated channel activities in guinea-pig gastric smooth muscle cells. Jpn J Physiol 46: 337–345, 1996. doi: 10.2170/jjphysiol.46.337. [DOI] [PubMed] [Google Scholar]
- 89.Mazzuoli-Weber G, Kugler EM, Buhler CI, Kreutz F, Demir IE, Ceyhan OG, Zeller F, Schemann M. Piezo proteins: incidence and abundance in the enteric nervous system. Is there a link with mechanosensitivity? Cell Tissue Res 375: 605–618, 2018. doi: 10.1007/s00441-018-2926-7. [DOI] [PubMed] [Google Scholar]
- 90.Kraichely RE, Strege PR, Sarr MG, Kendrick ML, Farrugia G. Lysophosphatidyl choline modulates mechanosensitive L-type Ca2+ current in circular smooth muscle cells from human jejunum. Am J Physiol Gastrointest Liver Physiol 296: G833–G839, 2009. doi: 10.1152/ajpgi.90610.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Gibbons SJ, Strege PR, Lei S, Roeder JL, Mazzone A, Ou Y, Rich A, Farrugia G. The alpha1H Ca2+ channel subunit is expressed in mouse jejunal interstitial cells of Cajal and myocytes. J Cell Mol Med 13: 4422–4431, 2009. doi: 10.1111/j.1582-4934.2008.00623.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Koh SD, Monaghan K, Ro S, Mason HS, Kenyon JL, Sanders KM. Novel voltage-dependent non-selective cation conductance in murine colonic myocytes. J Physiol 533: 341–355, 2001. doi: 10.1111/j.1469-7793.2001.0341a.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Xiong Z, Sperelakis N, Noffsinger A, Fenoglio-Preiser C. Ca2+ currents in human colonic smooth muscle cells. Am J Physiol Gastrointest Liver Physiol 269: G378–G385, 1995. doi: 10.1152/ajpgi.1995.269.3.G378. [DOI] [PubMed] [Google Scholar]
- 94.Kim YC, Koh SD, Sanders KM. Voltage-dependent inward currents of interstitial cells of Cajal from murine colon and small intestine. J Physiol 541: 797–810, 2002. doi: 10.1113/jphysiol.2002.018796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Strege PR, Sha L, Beyder A, Bernard CE, Perez-Reyes E, Evangelista S, Gibbons SJ, Szurszewski JH, Farrugia G. T-type Ca2+ channel modulation by otilonium bromide. Am J Physiol Gastrointest Liver Physiol 298: G706–G713, 2010. doi: 10.1152/ajpgi.00437.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Francois A, Schuetter N, Laffray S, Sanguesa J, Pizzoccaro A, Dubel S, Mantilleri A, Nargeot J, Noel J, Wood JN, Moqrich A, Pongs O, Bourinet E. The low-threshold calcium channel Cav3.2 determines low-threshold mechanoreceptor function. Cell Rep 10: 370–382, 2015. doi: 10.1016/j.celrep.2014.12.042. [DOI] [PubMed] [Google Scholar]
- 97.Gellens ME, George AL Jr, Chen LQ, Chahine M, Horn R, Barchi RL, Kallen RG. Primary structure and functional expression of the human cardiac tetrodotoxin-insensitive voltage-dependent sodium channel. Proc Natl Acad Sci USA 89: 554–558, 1992. doi: 10.1073/pnas.89.2.554. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Holm AN, Rich A, Miller SM, Strege P, Ou Y, Gibbons S, Sarr MG, Szurszewski JH, Rae JL, Farrugia G. Sodium current in human jejunal circular smooth muscle cells. Gastroenterology 122: 178–187, 2002. doi: 10.1053/gast.2002.30346. [DOI] [PubMed] [Google Scholar]
- 99.Strege PR, Mazzone A, Kraichely RE, Sha L, Holm AN, Ou Y, Lim I, Gibbons SJ, Sarr MG, Farrugia G. Species dependent expression of intestinal smooth muscle mechanosensitive sodium channels. Neurogastroenterol Motil 19: 135–143, 2007. doi: 10.1111/j.1365-2982.2006.00844.x. [DOI] [PubMed] [Google Scholar]
- 100.Strege PR, Ou Y, Sha L, Rich A, Gibbons SJ, Szurszewski JH, Sarr MG, Farrugia G. Sodium current in human intestinal interstitial cells of Cajal. Am J Physiol Gastrointest Liver Physiol 285: G1111–G1121, 2003. doi: 10.1152/ajpgi.00152.2003. [DOI] [PubMed] [Google Scholar]
- 101.Strege PR, Mazzone A, Bernard CE, Neshatian L, Gibbons SJ, Saito YA, Tester DJ, Calvert ML, Mayer EA, Chang L, Ackerman MJ, Beyder A, Farrugia G. Irritable bowel syndrome patients have SCN5A channelopathies that lead to decreased NaV1.5 current and mechanosensitivity. Am J Physiol Gastrointest Liver Physiol 314: G494–G503, 2018. doi: 10.1152/ajpgi.00016.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Strege PR, Mercado-Perez A, Mazzone A, Saito YA, Bernard CE, Farrugia G, Beyder A. SCN5A mutation G615E results in NaV1.5 voltage-gated sodium channels with normal voltage-dependent function yet loss of mechanosensitivity. Channels (Austin) 13: 287–298, 2019. doi: 10.1080/19336950.2019.1632670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Beyder A, Strege PR, Bernard C, Farrugia G. Membrane permeable local anesthetics modulate Na(V)1.5 mechanosensitivity. Channels (Austin) 6: 308–316, 2012. doi: 10.4161/chan.21202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Locke GR 3rd, Ackerman MJ, Zinsmeister AR, Thapa P, Farrugia G. Gastrointestinal symptoms in families of patients with an SCN5A-encoded cardiac channelopathy: evidence of an intestinal channelopathy. Am J Gastroenterol 101: 1299–1304, 2006. [DOI] [PubMed] [Google Scholar]
- 105.Zaytseva AK, Karpushev AV, Kiselev AM, Mikhaylov EN, Lebedev DS, Zhorov BS, Kostareva AA. Characterization of a novel SCN5A genetic variant A1294G associated with mixed clinical phenotype. Biochem Biophys Res Commun 516: 777–783, 2019. doi: 10.1016/j.bbrc.2019.06.080. [DOI] [PubMed] [Google Scholar]
- 106.Beyder A, Mazzone A, Strege PR, Tester DJ, Saito YA, Bernard CE, Enders FT, Ek WE, Schmidt PT, Dlugosz A, Lindberg G, Karling P, Ohlsson B, Gazouli M, Nardone G, Cuomo R, Usai-Satta P, Galeazzi F, Neri M, Portincasa P, Bellini M, Barbara G, Camilleri M, Locke GR, Talley NJ, D'Amato M, Ackerman MJ, Farrugia G. Loss-of-function of the voltage-gated sodium channel NaV1.5 (channelopathies) in patients with irritable bowel syndrome. Gastroenterology 146: 1659–1668, 2014. doi: 10.1053/j.gastro.2014.02.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Sanders KM. Regulation of smooth muscle excitation and contraction. Neurogastroenterol Motil 20: 39–53, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Hao J, Padilla F, Dandonneau M, Lavebratt C, Lesage F, Noel J, Delmas P. Kv1.1 channels act as mechanical brake in the senses of touch and pain. Neuron 77: 899–914, 2013. doi: 10.1016/j.neuron.2012.12.035. [DOI] [PubMed] [Google Scholar]
- 109.Khavandi K, Baylie RA, Sugden SA, Ahmed M, Csato V, Eaton P, Hill-Eubanks DC, Bonev AD, Nelson MT, Greenstein AS. Pressure-induced oxidative activation of PKG enables vasoregulation by Ca2+ sparks and BK channels. Sci Signal 9: ra100, 2016. doi: 10.1126/scisignal.aaf6625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Brohawn SG, Su Z, MacKinnon R. Mechanosensitivity is mediated directly by the lipid membrane in TRAAK and TREK1 K+ channels. Proc Natl Acad Sci USA 111: 3614–3619, 2014. doi: 10.1073/pnas.1320768111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Fukasaku M, Kimura J, Yamaguchi O. Swelling-activated and arachidonic acid-induced currents are TREK-1 in rat bladder smooth muscle cells. Fukushima J Med Sci 62: 18–26, 2016. doi: 10.5387/fms.2015-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Gu CX, Juranka PF, Morris CE. Stretch-activation and stretch-inactivation of Shaker-IR, a voltage-gated K+ channel. Biophys J 80: 2678–2693, 2001. doi: 10.1016/S0006-3495(01)76237-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Schmidt D, Del Marmol J, MacKinnon R. Mechanistic basis for low threshold mechanosensitivity in voltage-dependent K+ channels. Proc Natl Acad Sci USA 109: 10352–10357, , 2012. doi: 10.1073/pnas.1204700109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Knot HJ, Nelson MT. Regulation of membrane potential and diameter by voltage-dependent K+ channels in rabbit myogenic cerebral arteries. Am J Physiol Heart Circ Physiol 269: H348–H355, 1995. doi: 10.1152/ajpheart.1995.269.1.H348. [DOI] [PubMed] [Google Scholar]
- 115.Leblanc N, Wan X, Leung PM. Physiological role of Ca(2+)-activated and voltage-dependent K+ currents in rabbit coronary myocytes. Am J Physiol Cell Physiol 266: C1523–C1537, 1994. doi: 10.1152/ajpcell.1994.266.6.C1523. [DOI] [PubMed] [Google Scholar]
- 116.Farrugia G. Modulation of ionic currents in isolated canine and human jejunal circular smooth muscle cells by fluoxetine. Gastroenterology 110: 1438–1445, 1996. doi: 10.1053/gast.1996.v110.pm8613049. [DOI] [PubMed] [Google Scholar]
- 117.Liu DH, Huang X, Guo X, Meng XM, Wu YS, Lu HL, Zhang CM, Kim YC, Xu WX. Voltage dependent potassium channel remodeling in murine intestinal smooth muscle hypertrophy induced by partial obstruction. PLoS One 9: e86109, 2014. doi: 10.1371/journal.pone.0086109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Alessandri-Haber N, Dina OA, Chen X, Levine JD. TRPC1 and TRPC6 channels cooperate with TRPV4 to mediate mechanical hyperalgesia and nociceptor sensitization. J Neurosci 29: 6217–6228, 2009. doi: 10.1523/JNEUROSCI.0893-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Quick K, Zhao J, Eijkelkamp N, Linley JE, Rugiero F, Cox JJ, Raouf R, Gringhuis M, Sexton JE, Abramowitz J, Taylor R, Forge A, Ashmore J, Kirkwood N, Kros CJ, Richardson GP, Freichel M, Flockerzi V, Birnbaumer L, Wood JN. TRPC3 and TRPC6 are essential for normal mechanotransduction in subsets of sensory neurons and cochlear hair cells. Open Biol 2: 120068, 2012. doi: 10.1098/rsob.120068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Borbiro I, Badheka D, Rohacs T. Activation of TRPV1 channels inhibits mechanosensitive Piezo channel activity by depleting membrane phosphoinositides. Sci Signal 8: ra15, 2015. doi: 10.1126/scisignal.2005667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Kim EC, Choi SK, Lim M, Yeon SI, Lee YH. Role of endogenous ENaC and TRP channels in the myogenic response of rat posterior cerebral arteries. PLoS One 8: e84194, 2013. doi: 10.1371/journal.pone.0084194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Zeng WZ, Marshall KL, Min S, Daou I, Chapleau MW, Abboud FM, Liberles SD, Patapoutian A. PIEZOs mediate neuronal sensing of blood pressure and the baroreceptor reflex. Science 362: 464–467, 2018. doi: 10.1126/science.aau6324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Lee W, Leddy HA, Chen Y, Lee SH, Zelenski NA, McNulty AL, Wu J, Beicker KN, Coles J, Zauscher S, Grandl J, Sachs F, Guilak F, Liedtke WB. Synergy between Piezo1 and Piezo2 channels confers high-strain mechanosensitivity to articular cartilage. Proc Natl Acad Sci USA 111: E5114–E5122, 2014. doi: 10.1073/pnas.1414298111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Corey DP, Garcia-Anoveros J, Holt JR, Kwan KY, Lin SY, Vollrath MA, Amalfitano A, Cheung EL, Derfler BH, Duggan A, Geleoc GS, Gray PA, Hoffman MP, Rehm HL, Tamasauskas D, Zhang DS. TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells. Nature 432: 723–730, 2004. doi: 10.1038/nature03066. [DOI] [PubMed] [Google Scholar]
- 125.Maroto R, Raso A, Wood TG, Kurosky A, Martinac B, Hamill OP. TRPC1 forms the stretch-activated cation channel in vertebrate cells. Nat Cell Biol 7: 179–185, 2005. doi: 10.1038/ncb1218. [DOI] [PubMed] [Google Scholar]
- 126.Bautista DM, Jordt SE, Nikai T, Tsuruda PR, Read AJ, Poblete J, Yamoah EN, Basbaum AI, Julius D. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell 124: 1269–1282, 2006. doi: 10.1016/j.cell.2006.02.023. [DOI] [PubMed] [Google Scholar]
- 127.Dietrich A, Kalwa H, Storch U, Mederos y Schnitzler M, Salanova B, Pinkenburg O, Dubrovska G, Essin K, Gollasch M, Birnbaumer L, Gudermann T. Pressure-induced and store-operated cation influx in vascular smooth muscle cells is independent of TRPC1. Pflugers Arch 455: 465–477, 2007. doi: 10.1007/s00424-007-0314-3. [DOI] [PubMed] [Google Scholar]
- 128.Dietrich A, Mederos YSM, Gollasch M, Gross V, Storch U, Dubrovska G, Obst M, Yildirim E, Salanova B, Kalwa H, Essin K, Pinkenburg O, Luft FC, Gudermann T, Birnbaumer L. Increased vascular smooth muscle contractility in TRPC6-/- mice. Mol Cell Biol 25: 6980–6989, 2005. [Erratum in Mol Cell Biol 25: 11191, 2005]. doi: 10.1128/MCB.25.16.6980-6989.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Dobnikar L, Taylor AL, Chappell J, Oldach P, Harman JL, Oerton E, Dzierzak E, Bennett MR, Spivakov M, Jorgensen HF. Disease-relevant transcriptional signatures identified in individual smooth muscle cells from healthy mouse vessels. Nat Commun 9: 4567, 2018. [Erratum in Nat Commun 9: 5401, 2018]. doi: 10.1038/s41467-018-06891-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Pedroza AJ, Tashima Y, Shad R, Cheng P, Wirka R, Churovich S, Nakamura K, Yokoyama N, Cui JZ, Iosef C, Hiesinger W, Quertermous T, Fischbein MP. Single-cell transcriptomic profiling of vascular smooth muscle cell phenotype modulation in marfan syndrome aortic aneurysm. Arterioscler Thromb Vasc Biol 40: 2195–2211, 2020. doi: 10.1161/ATVBAHA.120.314670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Wirka RC, Wagh D, Paik DT, Pjanic M, Nguyen T, Miller CL, Kundu R, Nagao M, Coller J, Koyano TK, Fong R, Woo YJ, Liu B, Montgomery SB, Wu JC, Zhu K, Chang R, Alamprese M, Tallquist MD, Kim JB, Quertermous T. Atheroprotective roles of smooth muscle cell phenotypic modulation and the TCF21 disease gene as revealed by single-cell analysis. Nat Med 25: 1280–1289, 2019. doi: 10.1038/s41591-019-0512-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Cui Y, Zheng Y, Liu X, Yan L, Fan X, Yong J, Hu Y, Dong J, Li Q, Wu X, Gao S, Li J, Wen L, Qiao J, Tang F. Single-cell transcriptome analysis maps the developmental track of the human heart. Cell Rep 26: 1934–1950, 2019. doi: 10.1016/j.celrep.2019.01.079. [DOI] [PubMed] [Google Scholar]
- 133.Liu X, Chen W, Li W, Li Y, Priest JR, Zhou B, Wang J, Zhou Z. Single-cell RNA-Seq of the developing cardiac outflow tract reveals convergent development of the vascular smooth muscle cells. Cell Rep 28: 1346–1361, 2019. doi: 10.1016/j.celrep.2019.06.092. [DOI] [PubMed] [Google Scholar]
- 134.Avrahami D, Wang YJ, Schug J, Feleke E, Gao L, Liu C, Consortium H, Naji A, Glaser B, Kaestner KH; HPAP Consortium. Single-cell transcriptomics of human islet ontogeny defines the molecular basis of beta-cell dedifferentiation in T2D. Mol Metab 42: 101057, 2020. doi: 10.1016/j.molmet.2020.101057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Cadwell CR, Scala F, Fahey PG, Kobak D, Mulherkar S, Sinz FH, Papadopoulos S, Tan ZH, Johnsson P, Hartmanis L, Li S, Cotton RJ, Tolias KF, Sandberg R, Berens P, Jiang X, Tolias AS. Cell type composition and circuit organization of clonally related excitatory neurons in the juvenile mouse neocortex. Elife 9: e52951, 2020. doi: 10.7554/eLife.52951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Camunas-Soler J, Dai XQ, Hang Y, Bautista A, Lyon J, Suzuki K, Kim SK, Quake SR, MacDonald PE. Patch-Seq links single-cell transcriptomes to human islet dysfunction in diabetes. Cell Metab 31: 1017–1031, 2020. doi: 10.1016/j.cmet.2020.04.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Fuzik J, Zeisel A, Mate Z, Calvigioni D, Yanagawa Y, Szabo G, Linnarsson S, Harkany T. Integration of electrophysiological recordings with single-cell RNA-seq data identifies neuronal subtypes. Nat Biotechnol 34: 175–183, 2016. doi: 10.1038/nbt.3443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Moncada R, Barkley D, Wagner F, Chiodin M, Devlin JC, Baron M, Hajdu CH, Simeone DM, Yanai I. Integrating microarray-based spatial transcriptomics and single-cell RNA-seq reveals tissue architecture in pancreatic ductal adenocarcinomas. Nat Biotechnol 38: 333–342, 2020. [Erratum in Nat Biotechnol 38: 1476, 2020]. doi: 10.1038/s41587-019-0392-8. [DOI] [PubMed] [Google Scholar]
- 139.Faussone-Pellegrini MS, Pantalone D, Cortesini C. An ultrastructural study of the smooth muscle cells and nerve endings of the human stomach. J Submicrosc Cytol Pathol 21: 421–437, 1989. [PubMed] [Google Scholar]
- 140.Costa M, Dodds KN, Wiklendt L, Spencer NJ, Brookes SJ, Dinning PG. Neurogenic and myogenic motor activity in the colon of the guinea pig, mouse, rabbit, and rat. Am J Physiol Gastrointest Liver Physiol 305: G749–G759, 2013. doi: 10.1152/ajpgi.00227.2013. [DOI] [PubMed] [Google Scholar]
- 141.Tran L, Greenwood-Van MB. In a non-human primate model, aging disrupts the neural control of intestinal smooth muscle contractility in a region-specific manner. Neurogastroenterol Motil 26: 410–418, 2014. doi: 10.1111/nmo.12290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Bernard CE, Gibbons SJ, Gomez-Pinilla PJ, Lurken MS, Schmalz PF, Roeder JL, Linden D, Cima RR, Dozois EJ, Larson DW, Camilleri M, Zinsmeister AR, Pozo MJ, Hicks GA, Farrugia G. Effect of age on the enteric nervous system of the human colon. Neurogastroenterol Motil 21: 746–e746, 2009. doi: 10.1111/j.1365-2982.2008.01245.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
