Next Article in Journal
Cyclophilins and Their Functions in Abiotic Stress and Plant–Microbe Interactions
Next Article in Special Issue
Revisiting the Large-Conductance Calcium-Activated Potassium (BKCa) Channels in the Pulmonary Circulation
Previous Article in Journal
Multifunctionality of Nanosized Calcium Apatite Dual-Doped with Li+/Eu3+ Ions Related to Cell Culture Studies and Cytotoxicity Evaluation In Vitro
Previous Article in Special Issue
Restoration of Vitamin D Levels Improves Endothelial Function and Increases TASK-Like K+ Currents in Pulmonary Arterial Hypertension Associated with Vitamin D Deficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanosensitivity in Pulmonary Circulation: Pathophysiological Relevance of Stretch-Activated Channels in Pulmonary Hypertension

by
Solène Barbeau
1,2,
Guillaume Gilbert
3,
Guillaume Cardouat
1,2,
Isabelle Baudrimont
1,2,
Véronique Freund-Michel
1,2,
Christelle Guibert
1,2,
Roger Marthan
1,2,
Pierre Vacher
1,2,
Jean-François Quignard
1,2 and
Thomas Ducret
1,2,*
1
Centre de Recherche Cardio-Thoracique de Bordeaux, Univ. Bordeaux, U1045, F-33600 Pessac, France
2
INSERM, Centre de Recherche Cardio-Thoracique de Bordeaux, U1045, F-33600 Pessac, France
3
ORPHY, UFR Sciences et Techniques, University of Brest, EA 4324, F-29238 Brest, France
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(9), 1389; https://doi.org/10.3390/biom11091389
Submission received: 12 July 2021 / Revised: 17 September 2021 / Accepted: 19 September 2021 / Published: 21 September 2021

Abstract

:
A variety of cell types in pulmonary arteries (endothelial cells, fibroblasts, and smooth muscle cells) are continuously exposed to mechanical stimulations such as shear stress and pulsatile blood pressure, which are altered under conditions of pulmonary hypertension (PH). Most functions of such vascular cells (e.g., contraction, migration, proliferation, production of extracellular matrix proteins, etc.) depend on a key event, i.e., the increase in intracellular calcium concentration ([Ca2+]i) which results from an influx of extracellular Ca2+ and/or a release of intracellular stored Ca2+. Calcium entry from the extracellular space is a major step in the elevation of [Ca2+]i, involving a variety of plasmalemmal Ca2+ channels including the superfamily of stretch-activated channels (SAC). A common characteristic of SAC is that their gating depends on membrane stretch. In general, SAC are non-selective Ca2+-permeable cation channels, including proteins of the TRP (Transient Receptor Potential) and Piezo channel superfamily. As membrane mechano-transducers, SAC convert physical forces into biological signals and hence into a cell response. Consequently, SAC play a major role in pulmonary arterial calcium homeostasis and, thus, appear as potential novel drug targets for a better management of PH.

1. Introduction

Pulmonary arteries (PA) are continually subjected to mechanical forces exerted by circulating blood on the three-layered vessel wall (i.e., intima, media, and adventitia) (Figure 1). Among these forces, shear stress and pulsatile blood pressure are the most notable (for review see [1]). Shear stress, the frictional force generated by blood flow, is parallel to the vessel wall and mainly acts on pulmonary arterial endothelial cells (PAEC) which are longitudinally aligned and form the inner tunica of the vessels. Conversely, blood pressure is a perpendicular force that causes a circumferential and a longitudinal stretch to the vessel wall, proportionally to the cardiac output resistance. Blood pressure acts both on pulmonary arterial smooth muscle cells (PASMC) which align circumferentially and form the median layer, and, to a lesser extent, on PAEC.
Shear stress and blood pressure have opposite effects: the former, acting on PAEC, induces a vasodilatation, whereas the latter, acting on PASMC, exerts an intrinsic vasomotor mechanism termed “myogenic tone” [2]. However, these two phenomena are conveyed through an elevation of the intracellular Ca2+ concentration ([Ca2+]i) that involves the superfamily of stretch-activated channels (SAC) and acts as an intermediate to physiological responses. Indeed, shear stress increases the [Ca2+]i in PAEC leading to a vasodilatation via, particularly, nitric oxide (NO) production that, in turn, influences adjacent PASMC contraction. Intravascular pressure, acting on PASMC, is also directly transduced by SAC, which can serve as direct route for Ca2+ entry and/or cause membrane depolarization which secondarily activates voltage-dependent Ca2+ channels. The resulting [Ca2+]i increase subsequently activates PASMC contraction through activation of calmodulin and myosin light chain kinase.
Beside blood mechanical forces, the composition of the extracellular matrix (ECM) also contributes to arterial stiffness and may, per se, modulate mechanotransduction in the vessel wall. Indeed, there is a complex crosstalk between ECM and vascular cells: PASMC and pulmonary artery adventitial fibroblasts (PAAF) produce the ECM, which in turn interacts with cells through SAC and different receptors such as integrins, altering stiffness-dependent vascular cell activation [3].
Mechanical stretch affects all of the main cell types (i.e., PAEC, PASMC, and PAAF) of the vessel wall and adapts the vessel tone to pressure in physiological circumstances. An alteration of mechanical stimuli is observed under pathological conditions, such as pulmonary hypertension (PH). Indeed, PH is a group of multifactorial pathological cardiovascular disorders characterized by a progressive elevation of pulmonary arterial pressure leading to right ventricular hypertrophy, heart failure and, ultimately, to premature death [4]. The recent haemodynamic definition of PH states that pre-capillary pulmonary hypertension due to pulmonary vascular disease is diagnosed when mean pulmonary arterial pressure of over 20 mmHg is associated with abnormal pulmonary vascular resistance of 3 or more Wood Units [5]. PH is divided into five groups according to clinical, hemodynamic, etiological characteristics, and treatment strategy: idiopathic and heritable pulmonary arterial hypertension (PAH) (group 1), PH due to left heart disease (group 2), PH due to lung diseases and/or hypoxia (group 3), PH associated to chronic thromboembolism (group 4), and finally PH forms with unclear or multifaceted origins (group 5) [4].
This is accompanied by an alteration of mechanical stresses to the vessel wall: an increased myogenic tone [6], a reduced flow rate [7] that is responsible for a lower shear stress [8], and a progressive PA stiffening [3]. In case of endothelium injury, occurring in PH or PA catheterization, especially during the measurement of pulmonary arterial wedge pressure, PASMC can also be directly exposed to flow shear stress or elevated transmural flow shear stress [9].
Although many mechanosensitive cellular components and extracellular structures have been shown to contribute to mechanotransduction [10], in PA wall in particular, the scope of this review is deliberately limited to SAC activation.

2. Stretch-Activated Channels

Plasma membrane SAC are referred to as mechanotransducers since they convert physical forces into biological signals and hence into an adaptive cell response. In the late 1970s, their existence was postulated by Corey and Hudspeth [11], before being electrophysiologically characterized, a few years later, in embryonic chick skeletal muscle by Guharay and Sachs [12]. Since their original characterization, SAC have been identified in a variety of tissues and species including mammalian vascular cells (for review see [13,14,15]) and pulmonary vascular cells in particular [9,16,17,18,19,20,21,22,23,24,25,26,27].
A common characteristic of SAC is that their open probability increases with the applied pressure, i.e., that their gating depends on membrane stretch. In pulmonary vascular cells, Ca2+ permeable non-selective cationic SAC share similar general electrophysiological properties: I-V relationship of the evoked currents is almost linear with a reversal potential around 0 mV and a unitary conductance around 30 pS [17,18,19,22,24]. Unlike voltage-gated channels, very few pharmacological inhibitors are available for SAC. The only specific one is GsMTx-4, a peptide toxin isolated from tarantula venom [28], which has shown inhibitory effect on SAC in PAEC [26] and PASMC [19,20,21,22,24,25]. GsMTx-4 does not act on a standard ‘‘lock and key’’ model, but acts by its incorporation into boundary lipids surrounding the channel in a tension-dependent manner and thereby reducing the transfer of force from the lipid bilayer to the channel [28]. In the context of PH, most SAC inhibitors used in experimental studies are unselective blockers: amiloride and its analogs, aminoglycoside antibiotics such as streptomycin [19,21,26], disulfonic stilbene DIDS (4,4’-diisothiocyanatostilbene-2,2’-disulfonate) [17,22], and Gd3+ [17,19,21,25,26]. Their non-selectivity and potential side effects certainly explain the absence of clinical studies.
As reviewed in [1], three general models are proposed to account for the activation mechanisms of SAC: the “bilayer model”, the “tether model”, and the “secondary signal model” (Figure 2).
In the “bilayer model” (Figure 2a), the tension developed in the lipid bilayer itself is directly responsible for channel gating. Different mechanisms are also described for this model in which intrinsic mechanosensitivity of the channel depends on dimensional changes and/or hydrophobic mismatch (for review see [29]). This direct mechanical activation is supported by solubilization and liposome reconstitution experiments in which SAC activity was maintained, as well as the kinetic and the reversibility of the current elicited by stretch. In the “tether model” (Figure 2b), the force is transmitted to the channel via proteins located in the extracellular matrix or the cytoskeleton or both. Thus, as demonstrated by using agents that disrupt (such as cytochalasin and colchicine) or stabilize (such as phalloidin) cytoskeleton or extracellular matrix organization, SAC opening depends on the displacement of these proteins relative to the channel. In the “secondary signal model” (Figure 2c), a distant mechanical-sensitive protein (e.g., phospholipase A2 or phospholipase C (PLC)) can release a diffusible second messenger (such as arachidonic acid and diacylglycerol (DAG)) or activate a kinase (such as protein kinase C (PKC)), which, in turn can activate the channel. These different types of activation mechanisms could exist in pulmonary vascular cells. Indeed, in PAEC and PASMC, a direct stretch activation of SAC (“bilayer” and “tether” models) is supported by the fact that the stretch-induced current is turned on and off without a noticeable delay [17,19,21,26]. Furthermore, SAC currents were recorded in inside-out patch-clamp configuration, excluding the possible involvement of secondary messengers, at the exception of enzyme-linked channels [17]. In contrast, some studies suggest an indirect mechanical activation (“secondary signal model”) in which stretch leads to activation of PKC (via stimulation of PLC and production of DAG), which, in turn, activates SAC [17,24]. Moreover, these two activation modes most likely trigger converging pathways since numerous SAC are polymodal channels (see sections below).
At the cellular and tissue levels, several strategies have been developed to investigate SAC. Among cell-based assays (Figure 3), the most commonly used method is based on membrane deformation including amphipathic compounds, elastomeric pillars, elongation of flexible cell substrates, magnetic particles, osmotic challenges, patch membrane stretch, piezo-driven pressure, and shear stress (for review see [30] or [31]). These techniques have been applied to pulmonary vascular cells in vitro [9,16,17,18,19,20,21,22,23,24,25,26,27,32]. An alternative method consists in culturing cells in matrices of different stiffness in order to evaluate the impact of the environment matrix on the cells. To this end, PAAF were cultured in different stiffness polyacrylamide gels corresponding to a normotensive pulmonary artery (0.5 kPa) or mimicking mild-severe PH (3-10 kPa) [33]. Furthermore, the effect of stretch can also be studied in whole vessels for a more integrative approach using arteriography or myography (Figure 4). In the former, a microvessel is cannulated at both ends with glass micropipettes and placed in a microvascular flow system chamber, allowing intraluminal pressure increase. In the latter, one end of the segment is anchored to a stationary support and the other end is connected to a force-displacement transducer to monitor the vessel contraction under resting tension corresponding to an adapted transmural pressure. These protocols are commonly used to investigate myogenic tone [6,19,23] or basal stretch [22] in small isolated intralobar PA. Mechano-dependent contraction of vessels can also be induced by hypo-osmotic shocks [25].
As the increase in [Ca2+]i due to activation of Ca2+ channels plays a key role in fundamental cellular processes involved in the pathogenesis of PH such as contraction, migration, and proliferation, and thereby in vascular remodeling and hyperreactivity, PH is often referred to as a vascular channelopathy. Thus, in the pulmonary vasculature, which is continually subjected to mechanical forces, it is likely that facilitation of Ca2+ entry via SAC contributes to development and/or maintenance of the pathology. Indeed, in PH of group 3, hypoxia exposure could initiate, via the hypoxic pulmonary vasoconstriction phenomenon [34], vascular tone elevation and subsequent SAC activation, contributing to the early development of PH. In other PH groups, although not being the trigger, SAC activation following alteration of mechanical stresses to the vessel could contribute to the progression and/or maintenance of the pathology. Interestingly, SAC are more active in PASMC from chronically hypoxic (CH) or monocrotaline (MCT) rats, two animal models of PH, than in those from healthy rats [19,20,25]. Moreover, in the pulmonary vasculature of normoxic rats, myogenic tone, sensitive to SAC blockers, is minimal as compared to that in the pulmonary vasculature of CH rats [6,19,23]. In addition, hypoxia exposure induces immediate upregulation of mechanosensitive TRPV4 channel in rat PA [23]. Thus, SAC appear to play a major role in PA pathophysiological signaling. There is a growing body of evidence implicating members of the Transient Receptor Potential (TRP), Piezo and other families of ion channels in this mechanotransduction, which will be described in the following sections.

3. TRP Channels

TRP channels constitute a superfamily of non-selective cationic channels (Table 1) that exhibit a common structure composed of N- and C-terminal regions containing protein interaction motifs and six transmembrane domains (TM1-TM6). The putative ion conducting pore is located between the 5th and 6th TM domain. TRP channels are polymodal and, hence, can be modulated by a wide variety of stimuli such as cold, heat, pH, membrane potential, mechanical stress, neurohormonal signals, Ca2+ and Mg2+ ions, oxidative stress, intracellular ligands, as well as vasoactive factors such as angiotensin-II, arachidonic acid and its metabolites epoxyeicosatrienoic acids (EET), adenosine triphosphate (ATP), thrombin, endothelin-1 (ET-1), and serotonin (5-HT). According to their activation stimuli and the presence of regulatory domains on the cytosolic N- and C-termini, TRP superfamily is subdivided into seven main subfamilies: TRPC1-7 (Canonical), TRPV1-6 (Vanilloid), TRPM1-8 (Melastatin), TRPP1-4 (Polycystin), TRPML1-3 (Mucolipin), TRPA1 (Ankyrin), and TRPN (no mechanoreceptor potential C, NOMPC). Functional TRP are composed of four homo- or heteromultimeric subunits. These latter display conduction properties and regulatory mechanisms that are distinct from those of homomeric channels. TRP channels are expressed in a wide range of cell types and, as Ca2+ permeable channels, they are central actors in cellular Ca2+ homeostasis.
Although the relevance in mechanotransduction has not been examined for many of TRP channel isoforms playing important physiological functions in pulmonary vessels, their mechanosensitivity in other tissues has been demonstrated, suggesting their putative role as SAC in pulmonary vessels. Thus, we describe in this section their role in pulmonary vasculature and involvement in PH.

3.1. TRPC

Among TRPC1-7 channels, TRPC1, TRPC3, TRPC4, TRPC5, and TRPC6 are sensitive to mechanical stimulation [36]. Interestingly, these five isoforms are detected in PA from different species including rat or human, and some of them play a role in PH [67,68].

3.1.1. TRPC1

TRPC1 was the first member of this channel superfamily to be identified in drosophila, in 1995 [69]. This channel is expressed ubiquitously and can interact with some other TRP channels such as TRPC4, TRPC5, TRPC6, TRPP2, or TRPV4 [70]. TRPC1 is a mechanosensitive channel involved in store-operated calcium entry (SOCE) and receptor-operated calcium entry (ROCE) in various cell types playing a role in cell proliferation and migration [36,37].
In the pulmonary vasculature, previous studies have demonstrated that TRPC1 is expressed in PASMC and PAEC in rat, mouse, and human [68,71,72]. The role of TRPC1 in PA has mostly been studied under pathological condition of PH where it participates in the worsening phenotype of the disease. TRPC1 expression is upregulated at mRNA and protein levels in PA from PH animal models such as CH [68], MCT [73] or chronic ligation of the left main PA (thromboembolic PH) [74]. In PASMC from PH rat, TRPC1 is involved in basal increase in [Ca2+]i and SOCE currents [68,74,75]. This dysregulation of Ca2+ homeostasis increases PASMC proliferation, observed in PH [71], and promotes PA remodeling and altered contraction responses [73,76]. Moreover, Bone Morphogenetic Protein 4 (BMP4), which plays an important role in PA remodeling, increases SOCE currents through TRPC1 overexpression in PASMC. This process is dependent on p38-ERK1/2 MAPK pathway [77,78]. In a pre-clinical study using a CH-induced PH murine model, Sun et al. demonstrated that intratracheal TRPC1 siRNA delivery reduces the production of pro-fibrosis factors, increases endothelial nitric oxide synthase (eNOS) expression that favors vasodilatation, and decreases pro-apoptotic factors, overall attenuating PH phenotype [79]. In accordance with this study, pulmonary vasculature remodeling in response to CH is reduced in trpc1−/− mice [71]. Moreover, vasomotor tone and hypoxia-enhanced 5-HT vasoconstriction are also attenuated in these mice [76]. Furthermore, right ventricular hypertrophy and microvessels remodeling are suppressed in double knock out mice trpc1−/−/trpc6−/−. Finally, sildenafil, a phosphodiesterase-5 (PDE5) inhibitor currently used to treat PH, reduces TRPC1 expression, diminishing basal [Ca2+] and SOCE currents in PASMC and, consequently, the severity of PH in a CH rat model [75,80]. Overall, these findings indicate that TRPC1 plays a central role in Ca2+ homeostasis in PH and that regulating its expression and/or function can be used as potential therapeutic means in the disease.

3.1.2. TRPC3

TRPC3 is a mechanosensitive channel ubiquitously expressed, particularly in central nervous and cardiovascular systems. TRPC3 can interact with some other TRP channels such as TRPC1, TRPC6, or TRPC7 [70]. TRPC3 is involved in SOCE and ROCE playing various roles in cardiovascular diseases [38].
TRPC3 is expressed in human and rat PA [81,82], in both PASMC and PAEC [38]. Moreover, this channel is upregulated in PASMC from patients with pulmonary arterial hypertension (PAH), at mRNA and protein levels [50]. This increase is associated with an increased SOCE [83]. However, to our knowledge, few studies reported a distinct role for TRPC3 in pulmonary vasculature or in PH.

3.1.3. TRPC4

TRPC4 is a ubiquitously expressed channel involved in ROCE/SOCE currents. This channel is one of the main TRPC channels expressed in human, mouse, and rat PA. Although stretch-induced TRPC4 activation has not been shown to be involved in PH, it can be activated by a variety of stimuli including arachidonic acid, whose signaling pathway is implicated in mechanosensitivity [42,43].
In smooth muscle and endothelial cells, TRPC4 plays an important role in regulating microvascular permeability, agonist-dependent vasorelaxation, and gene transcription [67]. At cellular level, TRPC4 is involved in Ca2+ signaling, altering endothelial permeability in PAEC [44,47]. In pulmonary vasculature, TRPC4 promotes pulmonary arterial constriction and remodeling in rat SUGEN-Hypoxia PH group 1 model. Thus, trpc4−/− rats present attenuated PA remodeling, reduction of plexiform lesions, associated with an increase of survival [46,47].

3.1.4. TRPC5

Initially identified in the central nervous system, TRPC5 is a Ca2+ permeable non-selective channel expressed in a range of tissues and cell types. Its expression in the pulmonary vasculature is conflicting. In some studies, this channel was detected in PASMC and PAEC [48] whereas in others, it was not detected in distal PA or PASMC [49] or detected only in PASMC but not in PAEC [50]. Despite the fact that TRPC5 role is well documented in the systemic vasculature, particularly in pathophysiological regulation of vascular tone [60,84] or in angiogenesis [85], to our knowledge, no studies have reported a distinct role for this mechanosensitive channel in PA or PH.

3.1.5. TRPC6

TRPC6 is a mechanosensitive channel ubiquitously expressed in vasculature. There is some evidence that TRPC6 is involved in SOCE and ROCE, playing a role in numerous biological processes including cell proliferation, myogenic tone, and agonist-induced vasoconstriction modulation [37]. Moreover, TRPC6 is also directly activated by DAG.
In pulmonary vasculature, studies have demonstrated TRPC6 expression in PASMC and PAEC in mouse, rat, and human [86,87], in proximal and distal PA [48,49,82,88]. Moreover, TRPC6 is also present in pulmonary venous smooth muscle [89]. Interestingly, it plays various roles in PH. Firstly, TRPC6 expression is upregulated in human PA from PH patients [50] and PASMC from idiopathic pulmonary arterial hypertension (IPAH) patients [90], as well as in PH animal models such as CH [68] and MCT [91]. A polymorphism in TRPC6 is also associated with an increased risk to develop IPAH [92]. Secondly, in PASMC, TRPC6 promotes SOCE and ROCE currents, and PASMC proliferation [50,90,93]. Furthermore, TRPC6 has been implicated in the adaptive contraction response of pulmonary vasculature to alveolar hypoxia [87]. Thirdly, trpc6−/− mice present less pulmonary vasculature remodeling in response to CH associated with a decreased vasomotor tone and hypoxia-enhanced 5-HT vasoconstriction response, attenuating the pulmonary hypertensive phenotype [76,94]. In addition, bosentan, an endothelin receptor inhibitor currently used to treat PAH, reduces TRPC6 expression, inhibiting cell proliferation [90]. Finally, as for TRPC1, sildenafil reduces TRPC6-evoked SOCE current in PASMC and consequently the severity of PH in CH rat model [75,80].
Promising new inhibitor candidates have been recently studied for PH treatment. Indeed, in CH or MCT rat models, chrysin, a component of medicinal plants, inhibits PH-induced TRPC1/6 upregulation, modulates Ca2+ homeostasis, reduces PASMC proliferation, and reactive oxygen species production, attenuating the PH phenotype (right ventricular pressure and hypertrophy, and vascular remodeling) [95,96]. In the same way, in CH mouse model, chloroquine, used for antimalarial treatment in Plasmodium vivax malaria, reverses TRPC1/6 overexpression and decreases vascular remodeling [97]. Finally, in CH rat model, topotecan, a toposisomerase inhibitor already used for cancer therapy, inhibits CH-induced TRPC1/4/6 overexpression and reduces [Ca2+]i in PASMC, improving hemodynamic parameters and attenuating artery remodeling and right ventricle hypertrophy [98].

3.2. TRPV

Among TRPV channels, TRPV1 [99,100], TRPV2 [101], and TRPV4 [66,102,103] are known to be mechanosensitive [104], especially to an osmotic stimulation [105].

3.2.1. TRPV1

TRPV1 is expressed in PAEC and PASMC [21,106,107,108,109], as well as in rat distal pulmonary venous smooth muscle cells (SMC) [110]. This channel is permeable to Ca2+ and contributes, to a large extent, to many Ca2+-signaling pathways including proliferation and migration associated with PH [21,107].
To study the functional implication of TRPV1 in PASMC, a specific agonist, capsaicin, has been used to induce Ca2+ responses [21,64,107]. In these cells, TRPV1 activation promotes PASMC migration associated with specific changes in the cytoskeleton architecture such as reorganization of F-actin, tubulin, and intermediate filament networks as well as NFAT nuclear translocation [21,107]. Thus, TRPV1 could activate Ca2+/calcineurin/NFAT cascade signaling pathway known to play a crucial role in the physiopathology of PASMC, implicated in hypoxia-induced PA remodeling and proliferation [21]. TRPV1 mRNA and proteins are increased in human PASMC exposed to CH, an in vitro model of PH [108]. Consistent with these observations, both capsaicin-induced increase of [Ca2+]i and channel activity are enhanced in PASMC from IPAH patients [64]. On the contrary, inhibition of TRPV1 by means of antagonists or knockdown reduces human hypoxia-induced PH [108] and IPAH PASMC proliferation [64]. Another evidence of TRPV1 role in proliferation is the response of rat PASMC to acute exposure to silicium dioxide nanoparticles. After such an exposure, the increase of the [Ca2+]i and a proliferative response are greater in PASMC isolated from rats with PH than in those from control rats, and these responses are inhibited in the presence of TRPV1 antagonist [111]. Although TRPV1 expression is not modified in rat PASMC cultured under hypoxic conditions, it seems that hypoxia induces TRPV1 membrane translocation [21]. Likewise, hypoxia exposure does not potentiate the TRPV1-induced cytoskeleton reorganization but TRPV1 inhibition limits the effect of hypoxia on cytoskeletal changes suggesting an implication of TRPV1 in this remodeling [21]. More recently, a beneficial effect of TRPV1 in PH was demonstrated through TRPV1 involvement in human PA vasodilatation in response to cannabidiol [106]. This phenomenon could be explained by the expression of TRPV1 in PAEC. Given that most of the previous studies describing the implication of TRPV1 in proliferation and migration associated with PH were performed on isolated PASMC, the signaling pathways implicated in PAEC after TRPV1 activation could provide counterbalancing effects at the site of the entire vessel.

3.2.2. TRPV2

Although TRPV2 is expressed in PA, its function has been less studied. mRNA expression for TRPV2 is greater than for TRPV4 and TRPV1 in rat distal pulmonary vein SMC and this expression was confirmed by Western blot [110]. However, in PA, TRPV2 expression is lower than that of TRPV4 [108,109]. TRPV2 is notably expressed in human and rat PASMC [108,112]. TRPV2 current shows a weakly outward rectifying I/V curve similar to the one observed in response to thromboxane A2 analogue. This suggests that TRPV2 is a potential candidate for non-selective cationic response to thromboxane A2. It is noteworthy that thromboxane A2 analogue has been shown to play a critical role in rat hypoxic pulmonary vasoconstriction, interfering that TRPV2 could also be implicated [112].

3.2.3. TRPV4

TRPV4 is widely expressed in every layer of PA [109]. In PAEC, it plays a role in vasodilatation [113,114], whereas in PASMC it is involved in contraction, migration, and proliferation [9,65,108,115,116,117]; and in PAAF, it participates in remodeling processes [118]. All these mechanisms are either impaired (namely endothelial dysfunction) or enhanced (namely contraction and remodeling) in PH. Hence, TRPV4 activation in endothelium induces relaxation through NO-signaling and endothelium-derived hyperpolarizing factor [119]. Furthermore, TRPV4 interacts with eNOS in PAEC [114] and activation of unitary Ca2+ influx events through TRPV4 in these cells induces the release of NO, which initiates the guanylyl cyclase–protein kinase G pathway. Besides, guanylyl cyclase–protein kinase G signaling provides a negative feedback for TRPV4-eNOS coupling, thus regulating TRPV4 function [113]. Although TRPV4-induced Ca2+-signaling in PAEC is an important pathway for vasorelaxation, a dysregulation of this signaling is observed in PH. Indeed, in a rat model of PAH (SU5416 plus hypoxia) occurring through vascular endothelial growth factor receptor 2 (VEGFR2) inhibition, elevated production of mitochondrial-derived ROS induces a Ca2+ influx via TRPV4, which promotes enhanced migratory and proliferative responses in PAEC [120]. In PAAF, TRPV4 activation induces proliferation, migration, and extracellular matrix production (collagen I and fibronectin) thereby modulating the matrix stiffness, thus contributing to the adventitial remodeling occurring during PH [118]. PASMC express functional TRPV4 channels which are involved in increased [Ca2+]i, which triggers proliferation [111,115,117] and migration [107]. TRPV4 expression is increased in CH rat PASMC [23,65] associated with enhanced responses, providing evidence that TRPV4 is important in the development of the pathology. Indeed, the TRPV4 upregulation in PASMC is consistent with the higher basal [Ca2+]i observed after CH, an enhanced myogenic tone and 5-HT-induced vasoconstriction via the cytochrome P450 epoxygenase–epoxyeicosatrienoic acid pathway CH PA [23,121]. Nevertheless, whereas culturing PASMC in CH (1% O2 during 48 h) does not affect TRPV4 expression, the amplitude of the TRPV4-induced Ca2+ elevation and migratory response are more elevated [21], suggesting a direct role of hypoxia on TRPV4 channel function. The deletion of the trpv4 gene in mice has provided evidence of the implication of TRPV4 channel in the development of PH. Indeed, in CH trpv4−/− mice, all the features of PH are reduced, such as right ventricle heart hypertrophy, remodeling via muscularization of the arteries [23], PA pressure increase associated with hypoxic pulmonary vasoconstriction [122], and enhancement in 5-HT induced contraction [116]. Moreover, in PASMC from IPAH patients, TRPV4 expression and shear stress-induced Ca2+ response are increased, whereas this hypersensitivity to mechanical stimuli is reduced by TRPV4 inhibition using pharmacological antagonists or siRNA knockdown [9].

3.3. TRPM

Among TRPM channels, TRPM3, TRPM4, and TRPM7 are known to be mechanosensitive [104]. TRPM3 is sensitive to hypotonic cell swelling [51] whereas TRPM4 implication in mechanosensitivity seems to be indirect [55,123]. TRPM7 has been described as sensitive to both stretch and osmolarity [58,124].

3.3.1. TRPM3

TRPM3 mRNA expression in rat PA was shown by RT-qPCR [109]. Its expression is weaker than other TRPM channels such as TRPM8, TRPM4, and TRPM7 but higher than TRPM5 or TRPM6. However, no functional studies have been performed on TRPM3 in PA.

3.3.2. TRPM4

TRPM4 is widely expressed in various tissues including SMC of rat aorta, cerebral arteries, and PA. In PA, TRPM4 mRNA expression is, quantitatively, the second most important TRPM channel detected after TRPM8 [109]. TRPM4 is selective for monovalent cations and has a calcium-dependent activation (with an EC50 of 10 µM in native SMC) [125]. The function of TRPM4 in PA has not been studied but it was shown that this channel is involved in the regulation of cerebral artery SMC membrane potential and contractility. Indeed, its activation induces Na+ influx in SMC under physiological conditions and a down-regulation of its expression impairs pressure-induced SMC depolarization and vasoconstriction in isolated rat cerebral arteries [55]. Another study reported that trpm4−/− mice are hypertensive and present an increase in contractility of peripheral resistance vessels and cardiac function [126]. The role of TRPM4 in PH should thus be further investigated.

3.3.3. TRPM7

Similar to the other TRPM channels, TRPM7 expression in rat PA has been evaluated by RT-qPCR and its mRNA levels are similar to those of TRPM4 [109]. TRPM7 is a divalent cation permeable ion channel with a greater permeability to Mg2+ than Ca2+ [56] and has been shown to contribute to various phenomena involved in PH such as inflammation, fibrosis, cell migration, and cell proliferation [127]. Several studies show its implication in enhanced proliferation and vascular remodeling associated with PH. Song et al. described the role of TRPM7 in sensing fluid flow shear stress in human PASMC from IPAH or normal patients. They also showed that TRPM7, as well as TRPV4, are necessary for shear stress-mediated Ca2+ increase in PASMC and that TRPM7 is also responsible for Mg2+ influx. Moreover, in IPAH PASMC, shear stress-induced Mg2+ increase is greater than in normal PASMC. This augmentation is due to upregulated TRPM7 expression in these cells [9]. Thus, upregulated mechanosensitive channels such as TRPM7 in PASMC may contribute to the development of IPAH by inducing sustained pulmonary vasoconstriction leading to vascular remodeling. Confirming these findings, a study showed the therapeutic effect of Ophiocordyceps sinensis, an entomopathogenic fungus, known to exert antiproliferative and antiremodeling effects on PH through TRPM7 inhibition. They found that TRPM7 is expressed in fibroblast-like cells of the right ventricle and in PA, and that this expression is increased in PH rats. TRPM7 deletion in PH mice attenuates the signs of PH. Inhibition of TRPM7 activity via this fungus has an antiproliferative effect in IPAH PASMC, and vaso-dilates human PA [59]. However, an opposite effect was observed in another study where, in condition of PH, TRPM7 currents and free Mg2+ concentration were reduced in human or rat PASMC. There, TRPM7 inhibition increased proliferation and apoptosis resistance in PASMC through MEK/ERK pathway and exacerbated hypoxia-induced PH in vivo, whereas TRPM7 overexpression decreased PASMC proliferation and apoptosis resistance [128].

3.4. TRPA1

TRPA1 is also mechanosensitive and proposed as a candidate in mechanically gated hair cell transduction in the auditory response [129], or as shown in Merkel cells from hamster buccal mucosa [105]. However, it has not been studied in pulmonary vessels.

3.5. TRPP

TRPP channels consist of TRPP1, TRPP2, and TRPP3. Among TRPP family, TRPP1 can interact with Polycystin-1 [60]; TRPP2 along with Polycystin-1, TRPV4, and TRPC1, then functioning as mechanosensitive channels [61]. TRPP1 and TRPP2 are both expressed in vascular smooth muscle and EC from cerebral and mesenteric arteries playing a role in regulation of blood vessel function and myogenic tone [130,131]. However, roles of TRPP channels in pulmonary vasculature are not known and require further investigation.

4. Piezo Channels

Ten years ago, the discovery of Piezo channels as new SAC by Coste et al. [132] opened up new research opportunities in the field of mechanotransduction. Firstly described in pressure sensitive neurons, their involvement in vascular physiopathology has emerged. Piezo are encoded by two genes, Piezo1 and Piezo2 (formerly named Fam38A and Fam38B), coding for two different channels: Piezo1 and Piezo2. Piezo channels have the particularity to form a homotrimeric protein complex. Hence, Piezo channels possess a propeller-shaped architecture that comprises the central cap and three peripheral blades. Each monomer presents 38 transmembrane domains [133] instead of the most common six transmembrane domains as for TRP channels. Piezo channels are mainly permeable to Ca2+ but also to Na+, K+, and Mg2+, therefore being non-selective Ca2+ channels [134] (Table 2). They are activated by several mechanical stress forces such as physiological shear stress [135], cell wall tension or, in vitro, via a membrane crushing by a pipette, positive, and negative pressures [132,136] or a hypo-osmotic stress. Recently, ultrasound waves (500 kHz) have also been shown to activate those channels [137]. Piezo1 and 2 have similar properties but Piezo1 appears to be preferentially activated by a negative pressure while Piezo2 by a positive one. The channel sensitivity to pressure also differs: compared to Piezo1, Piezo2 channel is activated by higher pressures [138]. Furthermore, selective chemical agonists for Piezo1 channel have been described: Yoda1, which acts as a gating modulator by sensitizing the activation threshold of the channel and slows its inactivation kinetic [139,140]; Jedi1/2, which modulates the channel activation by potentializing its mechanosensitivity and slows down its inactivation kinetic [141]. Several common antagonists of SAC also antagonize Piezo1 channel such as the GsMTx-4 toxin [142,143], Gd3+ or ruthenium red [132,144]. However, some selective antagonists of the Yoda1-induced Piezo1 activation have been described such as Dooku1 (a Yoda1 analogue) [145] or Tubeimoside I (TBMS1) [146].
Piezo1 channels have been implicated in the pathogenesis of several vascular diseases. Deletion of Piezo1 gene in mouse leads to an aberrant vascular development resulting in an early embryonic death around day 10 [151]. In human, loss of function mutation of Piezo1 causes a congenital lymphatic dysplasia and patients are affected with a persistent lymphedema [152]. In systemic circulation, Piezo1 controls pressure or wall thickness [153,154]. Piezo1 channels display synergic or opposite functions in vessels, depending on the cell type where they are expressed. Endothelial Piezo1 channels are required for flow-induced vasodilatation due to eNOS activation and NO release [155]. Besides, the Ca2+ influx through these channels is not the only mechanism leading to NO formation as it also requires an ATP release through pannexin channels and subsequent P2Y2 receptor and endothelial NOS activation [155]. An alternative signaling pathway involves the release of adrenomedulin, which in turn, triggers NO release [156]. In uterine artery, as well as in the aorta, endothelial Piezo1 induces vessel relaxation [145,157]. By contrast, in mesenteric vessels, Piezo1 is responsible for flow-sensitive cationic ion influx in EC, which depolarizes the membrane. This depolarization spreads to the adjacent vascular SMC and activates voltage-gated Ca2+ channels causing vasoconstriction [158]. Retailleau et al. [153] reported that the increased cytosolic Ca2+ due to enhanced Piezo1 activation in vascular SMC influences both the diameter and wall thickness during hypertension in systemic circulation.
The presence of Piezo1 and 2 channels in the lung was described in the first report of Piezo as the tissue with the highest expression of these channels [132]. Others, as well as we, have reported the presence of Piezo1 in rat, mouse, or human PA [26,147,148,149]. Piezo2 is also present in human PA but its function remains unknown [147]. Human and mouse PAEC express Piezo1 channels [26]. The stimulation of this channel by a hypo-osmotic shock or by its selective agonist Yoda1 induces a large Ca2+ influx. The intracellular Ca2+ increase subsequently stimulates NO formation through eNOS, which brings about relaxation of mouse intrapulmonary arteries [26]. Unlike in the systemic circulation, Piezo1-induced NO formation does not require activation of ATP receptors. Selective deletion of Piezo1 in mouse EC inhibits Yoda-1-induced relaxation. Interestingly, endothelial deletion of Piezo1 increases α-adrenergic agonist-mediated contraction. Two hypotheses could explain this observation. On the one hand, phenylephrine-induced contraction of the vessel modifies its shape, which could lead to PAEC stretch, which, in turn, activates PAEC-Piezo1 channels. NO released by PAEC would then antagonize the contraction. On the other hand, phenylephrine could release factors derived from PASMC that may stimulate endothelial Piezo channels. In this connection, a crosstalk between SMC and endothelium has been previously described in pulmonary artery, where 5-HT stimulates both SMC contraction and NO release [159,160,161].
In CH mouse, Piezo1 still mediates arterial relaxation. Deletion of this channel in PAEC does not impair the development of the disease [26]. This result is supported by RT-qPCR measurements that show no variation in Piezo1 transcript in human PA from PH patients [147]. This finding is original as several other relaxation pathways are reduced during hypertension such as the acetylcholine pathway, which also requires NOS activation. As the Piezo1-mediated relaxation is still present in PH, activation of this channel could serve as therapeutic pathway.
In cultured human PASMC, Yoda1 induces a large intracellular Ca2+ increase due to activation of Piezo1 channels located at the plasma membrane and in intracellular organelles such as reticulum or mitochondria [149,162]. This Ca2+ rise through Piezo1 channels contributes to cell contraction and proliferation. In cultivated PASMC derived from IPAH patients, Piezo1 is upregulated and its activation triggers a higher Ca2+ influx (from both extracellular and intracellular Ca2+ stores) leading to an increased cell proliferation that could contribute to the pathophysiology of PH [149].
In conclusion, while endothelial Piezo1 function is not altered in mouse PA in PH, PASMC-Piezo1 channels are upregulated in human arteries [149], and could contribute to the development of the pathology. Furthermore, Piezo1 are also involved in several other pathophysiological processes in other tissues that share common pathological characteristics with PH. Indeed, activation of Piezo1 by hydrostatic pressure has been shown to activate inflammasome signaling pathways in macrophages [163] and in nucleus pulposus cells [164]. Turbulent flow induces Piezo1- and Gq/G11-mediated integrin activation, resulting in focal adhesion kinase-dependent NF-κB [165]. Those mechanisms could also contribute to PH and these pathways will also be interesting to explore in future studies. Moreover, the fact that Piezo1 activation is involved in free radical and NO production in PAEC [26] suggests that it could also contribute to the production of NOO, a toxic free radical that has been shown to be involved in PH [166]. Finally, Piezo1 induces cell migration and metastasis in several cancers [167]. Furthermore, Piezo1 localizes at focal adhesions to activate integrin-FAK signaling [168]. As Piezo1 is mainly expressed in the PAEC lamellipodia [151], we could postulate that Piezo1 could be involved in the remodeling of PA during PH.

5. Other Mechanosensitive Channels

In contrast to cationic depolarizing channels such as TRPV4 and Piezo, stretch also activates K+ hyperpolarizing channels (Table 3) such as two-pore domain K+ channels (K2P: TREK-1, TREK-2, TRAAK), ATP-dependent K+ channels (KATP) [169] or calcium-activated K+ channels (BKCa) [170]. However, stretch is not the only activator of these channels as they mainly respond to physiological chemical activators. Their activation, inducing a cellular hyperpolarization, acts as counterparts to TRPV4 or Piezo1 depolarization activity, enabling fine-tuning of mechanosensation, reducing cellular contraction in response to mechanical forces [171]. Potassium flux is also involved in cell proliferation and inflammation [172,173].
Whereas expression of K2P (TREK-1 especially) channels in PA has been confirmed [180,181], involvement of these channels in PH has not been described yet. KATP channels have been shown to be expressed in rabbit and human PASMC where they are involved in the modulation of the resting membrane potential [178,179]. Furthermore, the KATP opener, Iptakalim, has been shown to prevent pulmonary resistance vascular remodeling, cell proliferation, endothelial dysfunction [182,183], and PH [184]. Early studies showed that BKCa channels were activated by stretch and arachidonic acid in rabbit PASMC [175]. Besides, their activation prevents the development of PH in rat [176,185]. However, the mechanosensitive part of the function of these K+ channels has not been studied in detail in PA cells and further investigation needs to be done.

6. Conclusions

Mechanosensing is a key mechanism that intervenes in the regulation of the vascular tone. All the cellular layers of PA are able to sense mechanical forces, from PAEC to PASMC and PAAF. Each cell type responds differently and jointly to induce vasodilatation or vasocontraction, depending on the physiological needs and the sensed pressure. During PH, where mechanical stimulations in the PA are altered, modification of cellular processes occurs, mainly mediated through SAC (e.g., TRP and Piezo channels), see schematic Figure 5. Stretch-induced signaling pathways have been shown to disturb PAAF evoked-fibrosis, PAEC evoked-vasodilatation, PASMC evoked-contraction, and also their proliferation and migration. Altogether these alterations are involved in the pathophysiological phenotypes observed in PH. Future studies should aim to elucidate, in more details, mechanistic pathways that involve SAC in PA with the objective to target SAC in a pre-clinical approach. However, caution must be taken as mechanosensing is a key cellular mechanism that should not be annihilated at any cost in the whole organism and organs.

Author Contributions

Conceptualization, C.G., G.C., G.G., I.B., J.-F.Q., P.V., R.M., S.B., T.D., and V.F.-M.; drawing figures, G.G.; writing-original draft preparation, G.C., G.G., J.-F.Q., P.V., S.B., and T.D.; supervision, J.-F.Q. and T.D.; funding acquisition, C.G., I.B., J.-F.Q., P.V., R.M., T.D., and V.F.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the French National Research Agency (ANR), grant number ANR-19-CE44-0010-02 (the CANALBRET project); the New Aquitaine regional council (Région Nouvelle-Aquitaine), grant number AAPR2020I-2019-8151810 (the PHYSTRIG project); and the “Fondation pour la Recherche Médicale” (FRM), grant number DPC20171138954.

Acknowledgments

In memory of our regretted colleague Anne-Marie Vacher.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamashiro, Y.; Yanagisawa, H. The Molecular Mechanism of Mechanotransduction in Vascular Homeostasis and Disease. Clin. Sci. 2020, 134, 2399–2418. [Google Scholar] [CrossRef]
  2. Bayliss, W.M. On the Local Reactions of the Arterial Wall to Changes of Internal Pressure. J. Physiol. 1902, 28, 220–231. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, F.; Haeger, C.M.; Dieffenbach, P.B.; Sicard, D.; Chrobak, I.; Coronata, A.M.; Suarez Velandia, M.M.; Vitali, S.; Colas, R.A.; Norris, P.C.; et al. Distal Vessel Stiffening Is an Early and Pivotal Mechanobiological Regulator of Vascular Remodeling and Pulmonary Hypertension. JCI Insight 2016, 1, e86987. [Google Scholar] [CrossRef] [PubMed]
  4. Galie, N.; Humbert, M.; Vachiery, J.L.; Gibbs, S.; Lang, I.; Torbicki, A.; Simonneau, G.; Peacock, A.; Vonk Noordegraaf, A.; Beghetti, M.; et al. 2015 ESC/ERS Guidelines for the Diagnosis and Treatment of Pulmonary Hypertension: The Joint Task Force for the Diagnosis and Treatment of Pulmonary Hypertension of the European Society of Cardiology (ESC) and the European Respiratory Society (ERS): Endorsed by: Association for European Paediatric and Congenital Cardiology (AEPC), International Society for Heart and Lung Transplantation (ISHLT). Eur. Heart J. 2016, 37, 67–119. [Google Scholar] [CrossRef] [PubMed]
  5. Humbert, M.; Galie, N.; McLaughlin, V.V.; Rubin, L.J.; Simonneau, G. An Insider View on the World Symposium on Pulmonary Hypertension. Lancet Respir. Med. 2019, 7, 484–485. [Google Scholar] [CrossRef] [Green Version]
  6. Broughton, B.R.; Walker, B.R.; Resta, T.C. Chronic Hypoxia Induces Rho kinase-Dependent Myogenic Tone in Small Pulmonary Arteries. Am. J. Physiol. Lung Cell Mol. Physiol. 2008, 294, L797–L806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. De La Roque, E.D.; Thiaudiere, E.; Ducret, T.; Marthan, R.; Franconi, J.M.; Guibert, C.; Parzy, E. Effect of Chronic Hypoxia on Pulmonary Artery Blood Velocity in Rats as Assessed by Electrocardiography-Triggered Three-Dimensional Time-Resolved MR Angiography. NMR Biomed. 2011, 24, 225–230. [Google Scholar] [CrossRef]
  8. Wu, D.; Birukov, K. Endothelial Cell Mechano-Metabolomic Coupling to Disease States in the Lung Microvasculature. Front. Bioeng. Biotechnol. 2019, 7, 172. [Google Scholar] [CrossRef]
  9. Song, S.; Yamamura, A.; Yamamura, H.; Ayon, R.J.; Smith, K.A.; Tang, H.; Makino, A.; Yuan, J.X. Flow Shear Stress Enhances Intracellular Ca2+ Signaling in Pulmonary Artery Smooth Muscle Cells from Patients with Pulmonary Arterial Hypertension. Am. J. Physiol. Cell Physiol. 2014, 307, C373–C383. [Google Scholar] [CrossRef] [Green Version]
  10. Ingber, D.E. Cellular Mechanotransduction: Putting All the Pieces Together again. FASEB J. 2006, 20, 811–827. [Google Scholar] [CrossRef]
  11. Corey, D.P.; Hudspeth, A.J. Response Latency of Vertebrate Hair Cells. Biophys. J. 1979, 26, 499–506. [Google Scholar] [CrossRef] [Green Version]
  12. Guharay, F.; Sachs, F. Stretch-Activated Single Ion Channel Currents in Tissue-Cultured Embryonic Chick Skeletal Muscle. J. Physiol. 1984, 352, 685–701. [Google Scholar] [CrossRef]
  13. Folgering, J.H.; Sharif-Naeini, R.; Dedman, A.; Patel, A.; Delmas, P.; Honore, E. Molecular Basis of the Mammalian Pressure-Sensitive Ion Channels: Focus on Vascular Mechanotransduction. Prog. Biophys. Mol. Biol. 2008, 97, 180–195. [Google Scholar] [CrossRef] [PubMed]
  14. Beech, D.J.; Kalli, A.C. Force Sensing by Piezo Channels in Cardiovascular Health and Disease. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 2228–2239. [Google Scholar] [CrossRef] [PubMed]
  15. Douguet, D.; Patel, A.; Xu, A.; Vanhoutte, P.M.; Honore, E. Piezo Ion Channels in Cardiovascular Mechanobiology. Trends Pharmacol. Sci. 2019, 40, 956–970. [Google Scholar] [CrossRef] [Green Version]
  16. Bialecki, R.A.; Kulik, T.J.; Colucci, W.S. Stretching Increases Calcium Influx and Efflux in Cultured Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol. 1992, 263, L602–L606. [Google Scholar] [CrossRef] [PubMed]
  17. Park, K.S.; Kim, Y.; Lee, Y.H.; Earm, Y.E.; Ho, W.K. Mechanosensitive Cation Channels in Arterial Smooth Muscle Cells Are Activated by Diacylglycerol and Inhibited by Phospholipase C Inhibitor. Circ. Res. 2003, 93, 557–564. [Google Scholar] [CrossRef] [Green Version]
  18. Park, K.S.; Lee, H.A.; Earm, K.H.; Ko, J.H.; Earm, Y.E.; Kim, S.J. Differential Distribution of Mechanosensitive Nonselective Cation Channels in Systemic and Pulmonary Arterial Myocytes of Rabbits. J. Vasc. Res. 2006, 43, 347–354. [Google Scholar] [CrossRef]
  19. Ducret, T.; El Arrouchi, J.; Courtois, A.; Quignard, J.F.; Marthan, R.; Savineau, J.P. Stretch-Activated Channels in Pulmonary Arterial Smooth Muscle Cells from Normoxic and Chronically Hypoxic Rats. Cell Calcium 2010, 48, 251–259. [Google Scholar] [CrossRef]
  20. Gilbert, G.; Ducret, T.; Marthan, R.; Savineau, J.P.; Quignard, J.F. Stretch-Induced Ca2+ Signalling in Vascular Smooth Muscle Cells Depends on Ca2+ Store Segregation. Cardiovasc. Res. 2014, 103, 313–323. [Google Scholar] [CrossRef] [Green Version]
  21. Parpaite, T.; Cardouat, G.; Mauroux, M.; Gillibert-Duplantier, J.; Robillard, P.; Quignard, J.F.; Marthan, R.; Savineau, J.P.; Ducret, T. Effect of Hypoxia on TRPV1 and TRPV4 Channels in Rat Pulmonary Arterial Smooth Muscle Cells. Pflugers Arch. 2016, 468, 111–130. [Google Scholar] [CrossRef]
  22. Kim, H.J.; Yoo, H.Y.; Jang, J.H.; Lin, H.Y.; Seo, E.Y.; Zhang, Y.H.; Kim, S.J. Wall Stretch and Thromboxane A2 Activate NO Synthase (eNOS) in Pulmonary Arterial Smooth Muscle Cells via H2O2 and Akt-Dependent Phosphorylation. Pflugers Arch. 2016, 468, 705–716. [Google Scholar] [CrossRef]
  23. Yang, X.R.; Lin, A.H.; Hughes, J.M.; Flavahan, N.A.; Cao, Y.N.; Liedtke, W.; Sham, J.S. Upregulation of Osmo-Mechanosensitive TRPV4 Channel Facilitates Chronic Hypoxia-Induced Myogenic Tone and Pulmonary Hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2012, 302, L555–L568. [Google Scholar] [CrossRef] [Green Version]
  24. Lee, H.A.; Baek, E.B.; Park, K.S.; Jung, H.J.; Kim, J.I.; Kim, S.J.; Earm, Y.E. Mechanosensitive Nonselective Cation Channel Facilitation by Endothelin-1 Is Regulated by Protein Kinase C in Arterial Myocytes. Cardiovasc. Res. 2007, 76, 224–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Gilbert, G.; Ducret, T.; Savineau, J.P.; Marthan, R.; Quignard, J.F. Caveolae Are Involved in Mechanotransduction during Pulmonary Hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2016, 310, L1078–L1087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lhomme, A.; Gilbert, G.; Pele, T.; Deweirdt, J.; Henrion, D.; Baudrimont, I.; Campagnac, M.; Marthan, R.; Guibert, C.; Ducret, T.; et al. Stretch-Activated Piezo1 Channel in Endothelial Cells Relaxes Mouse Intrapulmonary Arteries. Am. J. Respir. Cell Mol. Biol. 2019, 60, 650–658. [Google Scholar] [CrossRef] [PubMed]
  27. Winston, F.K.; Thibault, L.E.; Macarak, E.J. An Analysis of the Time-Dependent Changes in Intracellular Calcium Concentration in Endothelial Cells in Culture Induced by Mechanical Stimulation. J. Biomech. Eng. 1993, 115, 160–168. [Google Scholar] [CrossRef] [PubMed]
  28. Gnanasambandam, R.; Ghatak, C.; Yasmann, A.; Nishizawa, K.; Sachs, F.; Ladokhin, A.S.; Sukharev, S.I.; Suchyna, T.M. GsMTx4: Mechanism of Inhibiting Mechanosensitive Ion Channels. Biophys. J. 2017, 112, 31–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kefauver, J.M.; Ward, A.B.; Patapoutian, A. Discoveries in Structure and Physiology of Mechanically Activated Ion Channels. Nature 2020, 587, 567–576. [Google Scholar] [CrossRef]
  30. Delmas, P.; Hao, J.; Rodat-Despoix, L. Molecular Mechanisms of Mechanotransduction in Mammalian Sensory Neurons. Nat. Rev. Neurosci. 2011, 12, 139–153. [Google Scholar] [CrossRef]
  31. Ranade, S.S.; Syeda, R.; Patapoutian, A. Mechanically Activated Ion Channels. Neuron 2015, 87, 1162–1179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Deweirdt, J.; Ducret, T.; Quignard, J.F.; Freund-Michel, V.; Lacomme, S.; Gontier, E.; Muller, B.; Marthan, R.; Guibert, C.; Baudrimont, I. Effects of FW2 Nanoparticles Toxicity in a New In Vitro Pulmonary Vascular Cells Model Mimicking Endothelial Dysfunction. Cardiovasc. Toxicol. 2021. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, A.; Cao, S.; Stowe, J.C.; Valdez-Jasso, D. Substrate Stiffness and Stretch Regulate Profibrotic Mechanosignaling in Pulmonary Arterial Adventitial Fibroblasts. Cells 2021, 10, 1000. [Google Scholar] [CrossRef] [PubMed]
  34. Dunham-Snary, K.J.; Wu, D.; Sykes, E.A.; Thakrar, A.; Parlow, L.R.G.; Mewburn, J.D.; Parlow, J.L.; Archer, S.L. Hypoxic Pulmonary Vasoconstriction: From Molecular Mechanisms to Medicine. Chest 2017, 151, 181–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Staruschenko, A.; Jeske, N.A.; Akopian, A.N. Contribution of TRPV1-TRPA1 Interaction to the Single Channel Properties of the TRPA1 Channel. J. Biol. Chem. 2010, 285, 15167–15177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Inoue, R.; Jian, Z.; Kawarabayashi, Y. Mechanosensitive TRP Channels in Cardiovascular Pathophysiology. Pharmacol. Ther. 2009, 123, 371–385. [Google Scholar] [CrossRef]
  37. Inoue, R.; Jensen, L.J.; Shi, J.; Morita, H.; Nishida, M.; Honda, A.; Ito, Y. Transient Receptor Potential Channels in Cardiovascular Function and Disease. Circ. Res. 2006, 99, 119–131. [Google Scholar] [CrossRef] [Green Version]
  38. Watanabe, H.; Murakami, M.; Ohba, T.; Takahashi, Y.; Ito, H. TRP Channel and Cardiovascular Disease. Pharmacol. Ther. 2008, 118, 337–351. [Google Scholar] [CrossRef]
  39. Morelli, M.B.; Amantini, C.; Liberati, S.; Santoni, M.; Nabissi, M. TRP Channels: New Potential Therapeutic Approaches in CNS Neuropathies. CNS Neurol. Disord. Drug Targets 2013, 12, 274–293. [Google Scholar] [CrossRef]
  40. Walker, R.L.; Koh, S.D.; Sergeant, G.P.; Sanders, K.M.; Horowitz, B. TRPC4 currents have properties similar to the pacemaker current in interstitial cells of Cajal. Am. J. Physiol. Cell Physiol. 2002, 283, C1637–C1645. [Google Scholar] [CrossRef] [Green Version]
  41. Freichel, M.; Tsvilovskyy, V.; Camacho-Londono, J.E. TRPC4- and TRPC4-Containing Channels. Handb. Exp. Pharmacol. 2014, 222, 85–128. [Google Scholar] [CrossRef]
  42. Vazquez, G.; Wedel, B.J.; Aziz, O.; Trebak, M.; Putney, J.W., Jr. The Mammalian TRPC Cation Channels. Biochim. Biophys. Acta 2004, 1742, 21–36. [Google Scholar] [CrossRef] [Green Version]
  43. Beech, D.J. Integration of Transient Receptor Potential Canonical Channels with Lipids. Acta Physiol. 2012, 204, 227–237. [Google Scholar] [CrossRef]
  44. Tiruppathi, C.; Freichel, M.; Vogel, S.M.; Paria, B.C.; Mehta, D.; Flockerzi, V.; Malik, A.B. Impairment of Store-Operated Ca2+ Entry in TRPC4−/− Mice Interferes with Increase in Lung Microvascular Permeability. Circ. Res. 2002, 91, 70–76. [Google Scholar] [CrossRef] [Green Version]
  45. Tiruppathi, C.; Ahmmed, G.U.; Vogel, S.M.; Malik, A.B. Ca2+ Signaling, TRP Channels, and Endothelial Permeability. Microcirculation 2006, 13, 693–708. [Google Scholar] [CrossRef] [PubMed]
  46. Alzoubi, A.; Almalouf, P.; Toba, M.; O’Neill, K.; Qian, X.; Francis, M.; Taylor, M.S.; Alexeyev, M.; McMurtry, I.F.; Oka, M.; et al. TRPC4 Inactivation Confers a Survival Benefit in Severe Pulmonary Arterial Hypertension. Am. J. Pathol. 2013, 183, 1779–1788. [Google Scholar] [CrossRef] [Green Version]
  47. Francis, M.; Xu, N.; Zhou, C.; Stevens, T. Transient Receptor Potential Channel 4 Encodes a Vascular Permeability Defect and High-Frequency Ca2+ Transients in Severe Pulmonary Arterial Hypertension. Am. J. Pathol. 2016, 186, 1701–1709. [Google Scholar] [CrossRef] [Green Version]
  48. McDaniel, S.S.; Platoshyn, O.; Wang, J.; Yu, Y.; Sweeney, M.; Krick, S.; Rubin, L.J.; Yuan, J.X. Capacitative Ca2+ Entry in Agonist-Induced Pulmonary Vasoconstriction. Am. J. Physiol. Lung Cell Mol. Physiol. 2001, 280, L870–L880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wang, J.; Shimoda, L.A.; Sylvester, J.T. Capacitative Calcium Entry and TRPC Channel Proteins Are Expressed in Rat Distal Pulmonary Arterial Smooth Muscle. Am. J. Physiol. Lung Cell Mol. Physiol. 2004, 286, L848–L858. [Google Scholar] [CrossRef] [PubMed]
  50. Yu, Y.; Fantozzi, I.; Remillard, C.V.; Landsberg, J.W.; Kunichika, N.; Platoshyn, O.; Tigno, D.D.; Thistlethwaite, P.A.; Rubin, L.J.; Yuan, J.X. Enhanced Expression of Transient Receptor Potential Channels in Idiopathic Pulmonary Arterial Hypertension. Proc. Natl. Acad. Sci. USA 2004, 101, 13861–13866. [Google Scholar] [CrossRef] [Green Version]
  51. Grimm, C.; Kraft, R.; Sauerbruch, S.; Schultz, G.; Harteneck, C. Molecular and Functional Characterization of the Melastatin-Related Cation Channel TRPM3. J. Biol. Chem. 2003, 278, 21493–21501. [Google Scholar] [CrossRef] [Green Version]
  52. Nilius, B.; Owsianik, G.; Voets, T.; Peters, J.A. Transient Receptor Potential Cation Channels in Disease. Physiol. Rev. 2007, 87, 165–217. [Google Scholar] [CrossRef] [Green Version]
  53. Launay, P.; Fleig, A.; Perraud, A.L.; Scharenberg, A.M.; Penner, R.; Kinet, J.P. TRPM4 Is a Ca2+-Activated Nonselective Cation Channel Mediating Cell Membrane Depolarization. Cell 2002, 109, 397–407. [Google Scholar] [CrossRef] [Green Version]
  54. Nilius, B.; Prenen, J.; Janssens, A.; Owsianik, G.; Wang, C.; Zhu, M.X.; Voets, T. The selectivity filter of the cation channel TRPM4. J. Biol. Chem. 2005, 280, 22899–22906. [Google Scholar] [CrossRef] [Green Version]
  55. Earley, S.; Waldron, B.J.; Brayden, J.E. Critical Role for Transient Receptor Potential Channel TRPM4 in Myogenic Constriction of Cerebral Arteries. Circ. Res. 2004, 95, 922–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Monteilh-Zoller, M.K.; Hermosura, M.C.; Nadler, M.J.; Scharenberg, A.M.; Penner, R.; Fleig, A. TRPM7 Provides an Ion Channel Mechanism for Cellular Entry of Trace Metal Ions. J. Gen. Physiol. 2003, 121, 49–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Runnels, L.W.; Yue, L.; Clapham, D.E. TRP-PLIK, a Bifunctional Protein with Kinase and Ion Channel Activities. Science 2001, 291, 1043–1047. [Google Scholar] [CrossRef] [PubMed]
  58. Numata, T.; Shimizu, T.; Okada, Y. TRPM7 Is a Stretch- and Swelling-Activated Cation Channel Involved in Volume Regulation in Human Epithelial Cells. Am. J. Physiol. Cell Physiol. 2007, 292, C460–C467. [Google Scholar] [CrossRef] [Green Version]
  59. Hiraishi, K.; Kurahara, L.H.; Feng, J.; Yamamura, A.; Cui, Y.; Yahiro, E.; Yokomise, H.; Go, T.; Ishikawa, K.; Yokota, N.; et al. Substantial Involvement of TRPM7 Inhibition in the Therapeutic Effect of Ophiocordyceps sinensis on Pulmonary Hypertension. Transl. Res. 2021, 233, 127–143. [Google Scholar] [CrossRef]
  60. Earley, S.; Brayden, J.E. Transient Receptor Potential Channels in the Vasculature. Physiol. Rev. 2015, 95, 645–690. [Google Scholar] [CrossRef] [Green Version]
  61. Du, J.; Fu, J.; Xia, X.M.; Shen, B. The Functions of TRPP2 in the Vascular System. Acta Pharmacol. Sin. 2016, 37, 13–18. [Google Scholar] [CrossRef] [Green Version]
  62. Guo, J.; Zhao, R.; Zhou, M.; Li, J.; Yao, X.; Du, J.; Chen, J.; Shen, B. TRPP2 and STIM1 form a Microdomain to Regulate Store-Operated Ca2+ Entry and Blood Vessel Tone. Cell Commun. Signal. 2020, 18, 138. [Google Scholar] [CrossRef] [PubMed]
  63. Sang, D.; Bai, S.; Yin, S.; Jiang, S.; Ye, L.; Hou, W.; Yao, Y.; Wang, H.; Shen, Y.; Shen, B.; et al. Role of TRPP2 in Mouse Airway Smooth Muscle Tension and Respiration. Am. J. Physiol. Lung Cell Mol. Physiol. 2019, 317, L466–L474. [Google Scholar] [CrossRef] [PubMed]
  64. Song, S.; Ayon, R.J.; Yamamura, A.; Yamamura, H.; Dash, S.; Babicheva, A.; Tang, H.; Sun, X.; Cordery, A.G.; Khalpey, Z.; et al. 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. 2017, 312, L309–L325. [Google Scholar] [CrossRef]
  65. Dahan, D.; Ducret, T.; Quignard, J.F.; Marthan, R.; Savineau, J.P.; Esteve, E. Implication of the Ryanodine Receptor in TRPV4-Induced Calcium Response in Pulmonary Arterial Smooth Muscle Cells from Normoxic and Chronically Hypoxic Rats. Am. J. Physiol. Lung Cell Mol. Physiol. 2012, 303, L824–L833. [Google Scholar] [CrossRef] [Green Version]
  66. Strotmann, R.; Harteneck, C.; Nunnenmacher, K.; Schultz, G.; Plant, T.D. OTRPC4, a Nonselective Cation Channel That Confers Sensitivity to Extracellular Osmolarity. Nat. Cell Biol. 2000, 2, 695–702. [Google Scholar] [CrossRef]
  67. Firth, A.L.; Remillard, C.V.; Yuan, J.X. TRP Channels in Hypertension. Biochim. Biophys. Acta 2007, 1772, 895–906. [Google Scholar] [CrossRef] [Green Version]
  68. Lin, M.J.; Leung, G.P.; Zhang, W.M.; Yang, X.R.; Yip, K.P.; Tse, C.M.; Sham, J.S. Chronic Hypoxia-Induced Upregulation of Store-Operated and Receptor-Operated Ca2+ Channels in Pulmonary Arterial Smooth Muscle Cells: A Novel Mechanism of Hypoxic Pulmonary Hypertension. Circ. Res. 2004, 95, 496–505. [Google Scholar] [CrossRef] [Green Version]
  69. Wes, P.D.; Chevesich, J.; Jeromin, A.; Rosenberg, C.; Stetten, G.; Montell, C. TRPC1, a Human Homolog of a Drosophila Store-Operated Channel. Proc. Natl. Acad. Sci. USA 1995, 92, 9652–9656. [Google Scholar] [CrossRef] [Green Version]
  70. Thakore, P.; Earley, S. Transient Receptor Potential Channels and Endothelial Cell Calcium Signaling. Compr. Physiol. 2019, 9, 1249–1277. [Google Scholar] [CrossRef] [PubMed]
  71. Malczyk, M.; Veith, C.; Fuchs, B.; Hofmann, K.; Storch, U.; Schermuly, R.T.; Witzenrath, M.; Ahlbrecht, K.; Fecher-Trost, C.; Flockerzi, V.; et al. Classical Transient Receptor Potential Channel 1 in hypoxia-Induced Pulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2013, 188, 1451–1459. [Google Scholar] [CrossRef] [PubMed]
  72. Kini, V.; Chavez, A.; Mehta, D. A New Role for PTEN in Regulating Transient Receptor Potential Canonical Channel 6-Mediated Ca2+ Entry, Endothelial Permeability, and Angiogenesis. J. Biol. Chem. 2010, 285, 33082–33091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Liu, X.R.; Zhang, M.F.; Yang, N.; Liu, Q.; Wang, R.X.; Cao, Y.N.; Yang, X.R.; Sham, J.S.; Lin, M.J. Enhanced Store-Operated Ca2+ Entry and TRPC Channel Expression in Pulmonary Arteries of Monocrotaline-Induced Pulmonary Hypertensive Rats. Am. J. Physiol. Cell Physiol. 2012, 302, C77–C87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yun, X.; Chen, Y.; Yang, K.; Wang, S.; Lu, W.; Wang, J. Upregulation of Canonical Transient Receptor Potential Channel in the Pulmonary Arterial Smooth Muscle of a Chronic Thromboembolic Pulmonary Hypertension Rat Model. Hypertens. Res. 2015, 38, 821–828. [Google Scholar] [CrossRef] [PubMed]
  75. Lu, W.; Ran, P.; Zhang, D.; Peng, G.; Li, B.; Zhong, N.; Wang, J. Sildenafil Inhibits Chronically Hypoxic Upregulation of Canonical Transient Receptor Potential Expression in Rat Pulmonary Arterial Smooth Muscle. Am. J. Physiol. Cell Physiol. 2010, 298, C114–C123. [Google Scholar] [CrossRef] [Green Version]
  76. Xia, Y.; Yang, X.R.; Fu, Z.; Paudel, O.; Abramowitz, J.; Birnbaumer, L.; Sham, J.S. Classical Transient Receptor Potential 1 and 6 Contribute to Hypoxic Pulmonary Hypertension through Differential Regulation of Pulmonary Vascular Functions. Hypertension 2014, 63, 173–180. [Google Scholar] [CrossRef] [Green Version]
  77. Zhang, Y.; Wang, Y.; Yang, K.; Tian, L.; Fu, X.; Wang, Y.; Sun, Y.; Jiang, Q.; Lu, W.; Wang, J. BMP4 Increases the Expression of TRPC and Basal [Ca2+]i via the p38MAPK and ERK1/2 Pathways Independent of BMPRII in PASMCs. PLoS ONE 2014, 9, e112695. [Google Scholar] [CrossRef] [Green Version]
  78. Li, X.; Lu, W.; Fu, X.; Zhang, Y.; Yang, K.; Zhong, N.; Ran, P.; Wang, J. BMP4 Increases Canonical Transient Receptor Potential Protein Expression by Activating p38 MAPK and ERK1/2 Signaling Pathways in Pulmonary Arterial Smooth Muscle Cells. Am. J. Respir. Cell Mol. Biol. 2013, 49, 212–220. [Google Scholar] [CrossRef] [Green Version]
  79. Sun, C.K.; Zhen, Y.Y.; Lu, H.I.; Sung, P.H.; Chang, L.T.; Tsai, T.H.; Sheu, J.J.; Chen, Y.L.; Chua, S.; Chang, H.W.; et al. Reducing TRPC1 Expression through Liposome-Mediated siRNA Delivery Markedly Attenuates Hypoxia-Induced Pulmonary Arterial Hypertension in a Murine Model. Stem Cells Int. 2014, 2014, 316214. [Google Scholar] [CrossRef]
  80. Wang, J.; Yang, K.; Xu, L.; Zhang, Y.; Lai, N.; Jiang, H.; Zhang, Y.; Zhong, N.; Ran, P.; Lu, W. Sildenafil Inhibits Hypoxia-Induced Transient Receptor Potential Canonical Protein Expression in Pulmonary Arterial Smooth Muscle via cGMP-PKG-PPARgamma Axis. Am. J. Respir. Cell Mol. Biol. 2013, 49, 231–240. [Google Scholar] [CrossRef] [Green Version]
  81. Gonzalez-Cobos, J.C.; Trebak, M. TRPC Channels in Smooth Muscle Cells. Front. Biosci. (Landmark Ed.) 2010, 15, 1023–1039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Rodat, L.; Savineau, J.P.; Marthan, R.; Guibert, C. Effect of Chronic Hypoxia on Voltage-Independent Calcium Influx Activated by 5-HT in Rat Intrapulmonary Arteries. Pflugers Arch. 2007, 454, 41–51. [Google Scholar] [CrossRef]
  83. Zhang, S.; Patel, H.H.; Murray, F.; Remillard, C.V.; Schach, C.; Thistlethwaite, P.A.; Insel, P.A.; Yuan, J.X. Pulmonary Artery Smooth Muscle Cells from Normal Subjects and IPAH Patients Show Divergent cAMP-Mediated Effects on TRPC Expression and Capacitative Ca2+ Entry. Am. J. Physiol. Lung Cell Mol. Physiol. 2007, 292, L1202–L1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Chen, X.; Yang, D.; Ma, S.; He, H.; Luo, Z.; Feng, X.; Cao, T.; Ma, L.; Yan, Z.; Liu, D.; et al. Increased rhythmicity in hypertensive arterial smooth muscle is linked to transient receptor potential canonical channels. J. Cell. Mol. Med. 2010, 14, 2483–2494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Negri, S.; Faris, P.; Berra-Romani, R.; Guerra, G.; Moccia, F. Endothelial Transient Receptor Potential Channels and Vascular Remodeling: Extracellular Ca2+ Entry for Angiogenesis, Arteriogenesis and Vasculogenesis. Front. Physiol. 2019, 10, 1618. [Google Scholar] [CrossRef] [Green Version]
  86. Singh, I.; Knezevic, N.; Ahmmed, G.U.; Kini, V.; Malik, A.B.; Mehta, D. Galphaq-TRPC6-Mediated Ca2+ Entry Induces RhoA Activation and Resultant Endothelial Cell Shape Change in Response to Thrombin. J. Biol. Chem. 2007, 282, 7833–7843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Malczyk, M.; Erb, A.; Veith, C.; Ghofrani, H.A.; Schermuly, R.T.; Gudermann, T.; Dietrich, A.; Weissmann, N.; Sydykov, A. The Role of Transient Receptor Potential Channel 6 Channels in the Pulmonary Vasculature. Front. Immunol. 2017, 8, 707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Ng, L.C.; Gurney, A.M. Store-Operated Channels Mediate Ca2+ Influx and Contraction in Rat Pulmonary Artery. Circ. Res. 2001, 89, 923–929. [Google Scholar] [CrossRef] [Green Version]
  89. Xu, L.; Chen, Y.; Yang, K.; Wang, Y.; Tian, L.; Zhang, J.; Wang, E.W.; Sun, D.; Lu, W.; Wang, J. Chronic Hypoxia Increases TRPC6 Expression and Basal Intracellular Ca2+ Concentration in Rat Distal Pulmonary venous Smooth Muscle. PLoS ONE 2014, 9, e112007. [Google Scholar] [CrossRef] [Green Version]
  90. Kunichika, N.; Landsberg, J.W.; Yu, Y.; Kunichika, H.; Thistlethwaite, P.A.; Rubin, L.J.; Yuan, J.X. Bosentan Inhibits Transient Receptor Potential Channel Expression in Pulmonary Vascular Myocytes. Am. J. Respir. Crit. Care Med. 2004, 170, 1101–1107. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, J.; Jiang, Q.; Wan, L.; Yang, K.; Zhang, Y.; Chen, Y.; Wang, E.; Lai, N.; Zhao, L.; Jiang, H.; et al. Sodium Tanshinone IIA Sulfonate Inhibits Canonical Transient Receptor Potential Expression in Pulmonary Arterial Smooth Muscle from Pulmonary hypertensive Rats. Am. J. Respir. Cell Mol. Biol. 2013, 48, 125–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Yu, Y.; Keller, S.H.; Remillard, C.V.; Safrina, O.; Nicholson, A.; Zhang, S.L.; Jiang, W.; Vangala, N.; Landsberg, J.W.; Wang, J.Y.; et al. A Functional Single-Nucleotide Polymorphism in the TRPC6 Gene Promoter Associated with Idiopathic Pulmonary Arterial Hypertension. Circulation 2009, 119, 2313–2322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Fernandez, R.A.; Wan, J.; Song, S.; Smith, K.A.; Gu, Y.; Tauseef, M.; Tang, H.; Makino, A.; Mehta, D.; Yuan, J.X. Upregulated Expression of STIM2, TRPC6, and Orai2 Contributes to the Transition of Pulmonary Arterial Smooth Muscle Cells from a Contractile to Proliferative Phenotype. Am. J. Physiol. Cell Physiol. 2015, 308, C581–C593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Smith, K.A.; Voiriot, G.; Tang, H.; Fraidenburg, D.R.; Song, S.; Yamamura, H.; Yamamura, A.; Guo, Q.; Wan, J.; Pohl, N.M.; et al. Notch Activation of Ca2+ Signaling in the Development of Hypoxic Pulmonary Vasoconstriction and Pulmonary Hypertension. Am. J. Respir. Cell Mol. Biol. 2015, 53, 355–367. [Google Scholar] [CrossRef] [Green Version]
  95. Dong, F.; Zhang, J.; Zhu, S.; Lan, T.; Yang, J.; Li, L. Chrysin Alleviates Chronic Hypoxia-Induced Pulmonary Hypertension by Reducing Intracellular Calcium Concentration in Pulmonary Arterial Smooth Muscle Cells. J. Cardiovasc. Pharmacol. 2019, 74, 426–435. [Google Scholar] [CrossRef]
  96. Dong, F.; Zhang, J.; Chen, X.; Zhang, S.; Zhu, L.; Peng, Y.; Guo, Z. Chrysin Alleviates Monocrotaline-Induced Pulmonary Hypertension in Rats through Regulation of Intracellular Calcium Homeostasis in Pulmonary Arterial Smooth Muscle Cells. J. Cardiovasc. Pharmacol. 2020, 75, 596–602. [Google Scholar] [CrossRef] [PubMed]
  97. Kaymak, E.; Akin, A.T.; Tufan, E.; Basaran, K.E.; Taheri, S.; Ozdamar, S.; Yakan, B. The Effect of Chloroquine on the TRPC1, TRPC6, and CaSR in the Pulmonary Artery Smooth Muscle Cells in Hypoxia-Induced Experimental Pulmonary Artery Hypertension. J. Biochem. Mol. Toxicol. 2021, 35, e22636. [Google Scholar] [CrossRef]
  98. Jiang, Y.; Zhou, Y.; Peng, G.; Liu, N.; Tian, H.; Pan, D.; Liu, L.; Yang, X.; Li, C.; Li, W.; et al. Topotecan Prevents Hypoxia-Induced Pulmonary Arterial Hypertension and Inhibits Hypoxia-Inducible Factor-1alpha and TRPC Channels. Int. J. Biochem. Cell Biol. 2018, 104, 161–170. [Google Scholar] [CrossRef]
  99. Ciura, S.; Bourque, C.W. Transient Receptor Potential Vanilloid 1 Is Required for Intrinsic Osmoreception in Organum Vasculosum Lamina Terminalis Neurons and for Normal Thirst Responses to Systemic Hyperosmolality. J. Neurosci. 2006, 26, 9069–9075. [Google Scholar] [CrossRef] [Green Version]
  100. Naeini, R.S.; Witty, M.F.; Seguela, P.; Bourque, C.W. An N-Terminal Variant of Trpv1 Channel Is Required for Osmosensory Transduction. Nat. Neurosci. 2006, 9, 93–98. [Google Scholar] [CrossRef]
  101. Muraki, K.; Iwata, Y.; Katanosaka, Y.; Ito, T.; Ohya, S.; Shigekawa, M.; Imaizumi, Y. TRPV2 Is a Component of Osmotically Sensitive Cation Channels in Murine Aortic Myocytes. Circ. Res. 2003, 93, 829–838. [Google Scholar] [CrossRef] [Green Version]
  102. Liedtke, W.; Choe, Y.; Marti-Renom, M.A.; Bell, A.M.; Denis, C.S.; Sali, A.; Hudspeth, A.J.; Friedman, J.M.; Heller, S. Vanilloid Receptor-Related Osmotically Activated Channel (VR-OAC), a Candidate Vertebrate Osmoreceptor. Cell 2000, 103, 525–535. [Google Scholar] [CrossRef] [Green Version]
  103. Nilius, B.; Prenen, J.; Wissenbach, U.; Bodding, M.; Droogmans, G. Differential Activation of the Volume-Sensitive Cation Channel TRP12 (OTRPC4) and Volume-Regulated Anion Currents in HEK-293 Cells. Pflugers Arch. 2001, 443, 227–233. [Google Scholar] [CrossRef] [PubMed]
  104. Sharif-Naeini, R.; Dedman, A.; Folgering, J.H.; Duprat, F.; Patel, A.; Nilius, B.; Honore, E. TRP Channels and Mechanosensory Transduction: Insights into the Arterial Myogenic Response. Pflugers Arch. 2008, 456, 529–540. [Google Scholar] [CrossRef] [Green Version]
  105. Soya, M.; Sato, M.; Sobhan, U.; Tsumura, M.; Ichinohe, T.; Tazaki, M.; Shibukawa, Y. Plasma membrane stretch activates transient receptor potential vanilloid and ankyrin channels in Merkel cells from hamster buccal mucosa. Cell Calcium 2014, 55, 208–218. [Google Scholar] [CrossRef]
  106. Baranowska-Kuczko, M.; Kozlowska, H.; Kloza, M.; Sadowska, O.; Kozlowski, M.; Kusaczuk, M.; Kasacka, I.; Malinowska, B. Vasodilatory Effects of Cannabidiol in Human Pulmonary and Rat Small Mesenteric Arteries: Modification by Hypertension and the Potential Pharmacological Opportunities. J. Hypertens. 2020, 38, 896–911. [Google Scholar] [CrossRef] [PubMed]
  107. Martin, E.; Dahan, D.; Cardouat, G.; Gillibert-Duplantier, J.; Marthan, R.; Savineau, J.P.; Ducret, T. Involvement of TRPV1 and TRPV4 Channels in Migration of Rat Pulmonary Arterial Smooth Muscle Cells. Pflugers Arch. 2012, 464, 261–272. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, Y.X.; Wang, J.; Wang, C.; Liu, J.; Shi, L.P.; Xu, M.; Wang, C. Functional Expression of Transient Receptor Potential Vanilloid-Related Channels in Chronically Hypoxic Human Pulmonary Arterial Smooth Muscle Cells. J. Membr. Biol. 2008, 223, 151–159. [Google Scholar] [CrossRef] [PubMed]
  109. Yang, X.R.; Lin, M.J.; McIntosh, L.S.; Sham, J.S. Functional Expression of Transient Receptor Potential Melastatin- and Vanilloid-Related Channels in Pulmonary Arterial and Aortic Smooth Muscle. Am. J. Physiol. Lung Cell Mol. Physiol. 2006, 290, L1267–L1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Peng, G.; Lu, W.; Li, X.; Chen, Y.; Zhong, N.; Ran, P.; Wang, J. Expression of Store-Operated Ca2+ Entry and Transient Receptor Potential Canonical and Vanilloid-Related Proteins in Rat Distal Pulmonary Venous Smooth Muscle. Am. J. Physiol. Lung Cell Mol. Physiol. 2010, 299, L621–L630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Dubes, V.; Parpaite, T.; Ducret, T.; Quignard, J.F.; Mornet, S.; Reinhardt, N.; Baudrimont, I.; Dubois, M.; Freund-Michel, V.; Marthan, R.; et al. Calcium Signalling Induced by In Vitro Exposure to Silicium Dioxide Nanoparticles in Rat Pulmonary Artery Smooth Muscle Cells. Toxicology 2017, 375, 37–47. [Google Scholar] [CrossRef]
  112. Yoo, H.Y.; Park, S.J.; Seo, E.Y.; Park, K.S.; Han, J.A.; Kim, K.S.; Shin, D.H.; Earm, Y.E.; Zhang, Y.H.; Kim, S.J. Role of Thromboxane A2-Activated Nonselective Cation Channels in Hypoxic Pulmonary Vasoconstriction of Rat. Am. J. Physiol. Cell Physiol. 2012, 302, C307–C317. [Google Scholar] [CrossRef] [Green Version]
  113. Marziano, C.; Hong, K.; Cope, E.L.; Kotlikoff, M.I.; Isakson, B.E.; Sonkusare, S.K. Nitric Oxide-Dependent Feedback Loop Regulates Transient Receptor Potential Vanilloid 4 (TRPV4) Channel Cooperativity and Endothelial Function in Small Pulmonary Arteries. J. Am. Heart Assoc. 2017, 6, e007157. [Google Scholar] [CrossRef] [Green Version]
  114. Ottolini, M.; Daneva, Z.; Chen, Y.L.; Cope, E.L.; Kasetti, R.B.; Zode, G.S.; Sonkusare, S.K. Mechanisms Underlying Selective Coupling of Endothelial Ca2+ Signals with eNOS vs. IK/SK Channels in Systemic and Pulmonary Arteries. J. Physiol. 2020, 598, 3577–3596. [Google Scholar] [CrossRef]
  115. Ducret, T.; Guibert, C.; Marthan, R.; Savineau, J.P. Serotonin-Induced Activation of TRPV4-like Current in Rat Intrapulmonary Arterial Smooth Muscle Cells. Cell Calcium 2008, 43, 315–323. [Google Scholar] [CrossRef]
  116. Xia, Y.; Fu, Z.; Hu, J.; Huang, C.; Paudel, O.; Cai, S.; Liedtke, W.; Sham, J.S. TRPV4 Channel Contributes to Serotonin-Induced Pulmonary Vasoconstriction and the Enhanced Vascular Reactivity in Chronic Hypoxic Pulmonary Hypertension. Am. J. Physiol. Cell Physiol. 2013, 305, C704–C715. [Google Scholar] [CrossRef] [Green Version]
  117. Yao, Z.; Wang, C. A Novel Mechanism of Sildenafil Improving the Excessive Proliferation and H2S Production in Pulmonary Arterial Smooth Muscle Cells. J. Cardiovasc. Pharmacol. 2019, 74, 355–363. [Google Scholar] [CrossRef]
  118. Cussac, L.A.; Cardouat, G.; Tiruchellvam Pillai, N.; Campagnac, M.; Robillard, P.; Montillaud, A.; Guibert, C.; Gailly, P.; Marthan, R.; Quignard, J.F.; et al. TRPV4 Channel Mediates Adventitial Fibroblast Activation and Adventitial Remodeling in Pulmonary Hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2020, 318, L135–L146. [Google Scholar] [CrossRef] [PubMed]
  119. Sukumaran, S.V.; Singh, T.U.; Parida, S.; Reddy, C.E.N.; Thangamalai, R.; Kandasamy, K.; Singh, V.; Mishra, S.K. TRPV4 Channel Activation Leads to Endothelium-Dependent Relaxation Mediated by Nitric Oxide and Endothelium-Derived Hyperpolarizing Factor in Rat Pulmonary Artery. Pharmacol. Res. 2013, 78, 18–27. [Google Scholar] [CrossRef] [PubMed]
  120. Suresh, K.; Servinsky, L.; Jiang, H.; Bigham, Z.; Yun, X.; Kliment, C.; Huetsch, J.; Damarla, M.; Shimoda, L.A. Reactive Oxygen Species Induced Ca2+ Influx via TRPV4 and Microvascular Endothelial Dysfunction in the SU5416/Hypoxia Model of Pulmonary Arterial Hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2018, 314, L893–L907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Xia, Y.; Xia, L.; Jin, Z.; Jin, R.; Paudel, O.; Sham, J.S.K. Cytochrome P450 Epoxygenase-Dependent Activation of TRPV4 Channel Participates in Enhanced Serotonin-Induced Pulmonary Vasoconstriction in Chronic Hypoxic Pulmonary Hypertension. Anal. Cell. Pathol. 2020, 2020, 8927381. [Google Scholar] [CrossRef] [Green Version]
  122. Goldenberg, N.M.; Wang, L.; Ranke, H.; Liedtke, W.; Tabuchi, A.; Kuebler, W.M. TRPV4 Is Required for Hypoxic Pulmonary Vasoconstriction. Anesthesiology 2015, 122, 1338–1348. [Google Scholar] [CrossRef] [PubMed]
  123. Morita, H.; Honda, A.; Inoue, R.; Ito, Y.; Abe, K.; Nelson, M.T.; Brayden, J.E. Membrane Stretch-Induced Activation of a TRPM4-like Nonselective Cation Channel in Cerebral Artery Myocytes. J. Pharmacol. Sci. 2007, 103, 417–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Numata, T.; Shimizu, T.; Okada, Y. Direct mechano-stress sensitivity of TRPM7 channel. Cell. Physiol. Biochem. 2007, 19, 1–8. [Google Scholar] [CrossRef] [PubMed]
  125. Earley, S. TRPM4 Channels in Smooth Muscle Function. Pflugers Arch. 2013, 465, 1223–1231. [Google Scholar] [CrossRef] [Green Version]
  126. Mathar, I.; Vennekens, R.; Meissner, M.; Kees, F.; Van der Mieren, G.; Londono, J.E.C.; Uhl, S.; Voets, T.; Hummel, B.; van den Bergh, A.; et al. Increased Catecholamine Secretion Contributes to Hypertension in TRPM4-Deficient Mice. J. Clin. Investig. 2010, 120, 3267–3279. [Google Scholar] [CrossRef] [Green Version]
  127. Touyz, R.M. Transient Receptor Potential Melastatin 6 and 7 Channels, Magnesium Transport, and Vascular Biology: Implications in Hypertension. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H1103–H1118. [Google Scholar] [CrossRef] [Green Version]
  128. Xing, J.; Wang, M.; Hong, J.; Gao, Y.; Liu, Y.; Gu, H.; Dong, J.; Li, L. TRPM7 channel inhibition exacerbates pulmonary arterial hypertension through MEK/ERK pathway. Aging 2019, 11, 4050–4065. [Google Scholar] [CrossRef]
  129. Corey, D.P.; Garcia-Anoveros, J.; Holt, J.R.; Kwan, K.Y.; Lin, S.Y.; Vollrath, M.A.; Amalfitano, A.; Cheung, E.L.; Derfler, B.H.; Duggan, A.; et al. TRPA1 Is a Candidate for the Mechanosensitive Transduction Channel of Vertebrate Hair Cells. Nature 2004, 432, 723–730. [Google Scholar] [CrossRef]
  130. Sharif-Naeini, R.; Folgering, J.H.; Bichet, D.; Duprat, F.; Lauritzen, I.; Arhatte, M.; Jodar, M.; Dedman, A.; Chatelain, F.C.; Schulte, U.; et al. Polycystin-1 and -2 Dosage Regulates Pressure Sensing. Cell 2009, 139, 587–596. [Google Scholar] [CrossRef] [Green Version]
  131. Narayanan, D.; Bulley, S.; Leo, M.D.; Burris, S.K.; Gabrick, K.S.; Boop, F.A.; Jaggar, J.H. Smooth Muscle Cell Transient Receptor Potential Polycystin-2 (TRPP2) Channels Contribute to the Myogenic Response in Cerebral Arteries. J. Physiol. 2013, 591, 5031–5046. [Google Scholar] [CrossRef]
  132. Coste, B.; Mathur, J.; Schmidt, M.; Earley, T.J.; Ranade, S.; Petrus, M.J.; Dubin, A.E.; Patapoutian, A. Piezo1 and Piezo2 Are Essential Components of Distinct Mechanically Activated Cation Channels. Science 2010, 330, 55–60. [Google Scholar] [CrossRef] [Green Version]
  133. Zhao, Q.; Zhou, H.; Chi, S.; Wang, Y.; Wang, J.; Geng, J.; Wu, K.; Liu, W.; Zhang, T.; Dong, M.Q.; et al. Structure and Mechanogating Mechanism of the Piezo1 Channel. Nature 2018, 554, 487–492. [Google Scholar] [CrossRef] [PubMed]
  134. Gnanasambandam, R.; Bae, C.; Gottlieb, P.A.; Sachs, F. Ionic Selectivity and Permeation Properties of Human PIEZO1 Channels. PLoS ONE 2015, 10, e0125503. [Google Scholar] [CrossRef] [Green Version]
  135. Morley, L.C.; Shi, J.; Gaunt, H.J.; Hyman, A.J.; Webster, P.J.; Williams, C.; Forbes, K.; Walker, J.J.; Simpson, N.A.B.; Beech, D.J. Piezo1 Channels Are Mechanosensors in Human Fetoplacental Endothelial Cells. Mol. Hum. Reprod. 2018, 24, 510–520. [Google Scholar] [CrossRef] [Green Version]
  136. Shin, K.C.; Park, H.J.; Kim, J.G.; Lee, I.H.; Cho, H.; Park, C.; Sung, T.S.; Koh, S.D.; Park, S.W.; Bae, Y.M. The Piezo2 Ion Channel Is Mechanically Activated by Low-Threshold Positive Pressure. Sci. Rep. 2019, 9, 6446. [Google Scholar] [CrossRef] [Green Version]
  137. Qiu, Z.; Guo, J.; Kala, S.; Zhu, J.; Xian, Q.; Qiu, W.; Li, G.; Zhu, T.; Meng, L.; Zhang, R.; et al. The Mechanosensitive Ion Channel Piezo1 Significantly Mediates In Vitro Ultrasonic Stimulation of Neurons. iScience 2019, 21, 448–457. [Google Scholar] [CrossRef] [PubMed]
  138. Du, G.; Li, L.; Zhang, X.; Liu, J.; Hao, J.; Zhu, J.; Wu, H.; Chen, W.; Zhang, Q. Roles of TRPV4 and Piezo Channels in Stretch-Evoked Ca2+ Response in Chondrocytes. Exp. Biol. Med. 2020, 245, 180–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Syeda, R.; Xu, J.; Dubin, A.E.; Coste, B.; Mathur, J.; Huynh, T.; Matzen, J.; Lao, J.; Tully, D.C.; Engels, I.H.; et al. Chemical Activation of the Mechanotransduction Channel Piezo1. eLife 2015, 4, e07369. [Google Scholar] [CrossRef]
  140. Lacroix, J.J.; Botello-Smith, W.M.; Luo, Y. Probing the Gating Mechanism of the Mechanosensitive Channel Piezo1 with the Small Molecule Yoda1. Nat. Commun. 2018, 9, 2029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Wang, Y.; Chi, S.; Guo, H.; Li, G.; Wang, L.; Zhao, Q.; Rao, Y.; Zu, L.; He, W.; Xiao, B. A Lever-like Transduction Pathway for Long-Distance Chemical- and Mechano-Gating of the Mechanosensitive Piezo1 Channel. Nat. Commun. 2018, 9, 1300. [Google Scholar] [CrossRef] [PubMed]
  142. Bae, C.; Sachs, F.; Gottlieb, P.A. The Mechanosensitive Ion Channel Piezo1 Is Inhibited by the Peptide GsMTx4. Biochemistry 2011, 50, 6295–6300. [Google Scholar] [CrossRef] [Green Version]
  143. Alcaino, C.; Knutson, K.; Gottlieb, P.A.; Farrugia, G.; Beyder, A. Mechanosensitive ion Channel Piezo2 Is Inhibited by D-GsMTx4. Channels 2017, 11, 245–253. [Google Scholar] [CrossRef] [Green Version]
  144. Coste, B.; Xiao, B.; Santos, J.S.; Syeda, R.; Grandl, J.; Spencer, K.S.; Kim, S.E.; Schmidt, M.; Mathur, J.; Dubin, A.E.; et al. Piezo Proteins Are Pore-Forming Subunits of Mechanically Activated Channels. Nature 2012, 483, 176–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Evans, E.L.; Cuthbertson, K.; Endesh, N.; Rode, B.; Blythe, N.M.; Hyman, A.J.; Hall, S.J.; Gaunt, H.J.; Ludlow, M.J.; Foster, R.; et al. Yoda1 Analogue (Dooku1) Which Antagonizes Yoda1-Evoked Activation of Piezo1 and Aortic Relaxation. Br. J. Pharmacol. 2018, 175, 1744–1759. [Google Scholar] [CrossRef]
  146. Liu, S.; Pan, X.; Cheng, W.; Deng, B.; He, Y.; Zhang, L.; Ning, Y.; Li, J. Tubeimoside I Antagonizes Yoda1-Evoked Piezo1 Channel Activation. Front. Pharmacol. 2020, 11, 768. [Google Scholar] [CrossRef] [PubMed]
  147. Lambert, M.; Capuano, V.; Olschewski, A.; Sabourin, J.; Nagaraj, C.; Girerd, B.; Weatherald, J.; Humbert, M.; Antigny, F. Ion Channels in Pulmonary Hypertension: A Therapeutic Interest? Int. J. Mol. Sci. 2018, 19, 3162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Jiang, L.; Zhang, Y.; Lu, D.; Huang, T.; Yan, K.; Yang, W.; Gao, J. Mechanosensitive Piezo1 Channel Activation Promotes Ventilator-Induced Lung Injury via Disruption of Endothelial Junctions in ARDS Rats. Biochem. Biophys. Res. Commun. 2021, 556, 79–86. [Google Scholar] [CrossRef]
  149. Liao, J.; Lu, W.; Chen, Y.; Duan, X.; Zhang, C.; Luo, X.; Lin, Z.; Chen, J.; Liu, S.; Yan, H.; et al. Upregulation of Piezo1 (Piezo Type Mechanosensitive Ion Channel Component 1) Enhances the Intracellular Free Calcium in Pulmonary Arterial Smooth Muscle Cells from Idiopathic Pulmonary Arterial Hypertension Patients. Hypertension 2021, 77, 1974–1989. [Google Scholar] [CrossRef]
  150. Bhattacharya, J.; Hough, R.F. Piezo1 in the Lung: At Last! Am. J. Respir. Cell Mol. Biol. 2019, 60, 609–610. [Google Scholar] [CrossRef]
  151. Li, J.; Hou, B.; Tumova, S.; Muraki, K.; Bruns, A.; Ludlow, M.J.; Sedo, A.; Hyman, A.J.; McKeown, L.; Young, R.S.; et al. Piezo1 Integration of Vascular Architecture with Physiological Force. Nature 2014, 515, 279–282. [Google Scholar] [CrossRef] [PubMed]
  152. Lukacs, V.; Mathur, J.; Mao, R.; Bayrak-Toydemir, P.; Procter, M.; Cahalan, S.M.; Kim, H.J.; Bandell, M.; Longo, N.; Day, R.W.; et al. Impaired PIEZO1 Function in Patients with a Novel Autosomal Recessive Congenital Lymphatic Dysplasia. Nat. Commun. 2015, 6, 8329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Retailleau, K.; Duprat, F.; Arhatte, M.; Ranade, S.S.; Peyronnet, R.; Martins, J.R.; Jodar, M.; Moro, C.; Offermanns, S.; Feng, Y.; et al. Piezo1 in Smooth Muscle Cells Is Involved in Hypertension-Dependent Arterial Remodeling. Cell Rep. 2015, 13, 1161–1171. [Google Scholar] [CrossRef] [Green Version]
  154. Allison, S.J. Hypertension: Mechanosensation by PIEZO1 in Blood Pressure Control. Nat. Rev. Nephrol. 2017, 13, 3. [Google Scholar] [CrossRef] [PubMed]
  155. Wang, S.; Chennupati, R.; Kaur, H.; Iring, A.; Wettschureck, N.; Offermanns, S. Endothelial Cation Channel PIEZO1 Controls Blood Pressure by Mediating Flow-Induced ATP Release. J. Clin. Investig. 2016, 126, 4527–4536. [Google Scholar] [CrossRef] [PubMed]
  156. Iring, A.; Jin, Y.J.; Albarran-Juarez, J.; Siragusa, M.; Wang, S.; Dancs, P.T.; Nakayama, A.; Tonack, S.; Chen, M.; Kunne, C.; et al. Shear Stress-Induced Endothelial Adrenomedullin Signaling Regulates Vascular Tone and Blood Pressure. J. Clin. Investig. 2019, 129, 2775–2791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. John, L.; Ko, N.L.; Gokin, A.; Gokina, N.; Mandala, M.; Osol, G. The Piezo1 Cation Channel Mediates Uterine Artery Shear Stress Mechanotransduction and Vasodilation during Rat Pregnancy. Am. J. Physiol. Heart Circ. Physiol. 2018, 315, H1019–H1026. [Google Scholar] [CrossRef] [PubMed]
  158. Rode, B.; Shi, J.; Endesh, N.; Drinkhill, M.J.; Webster, P.J.; Lotteau, S.J.; Bailey, M.A.; Yuldasheva, N.Y.; Ludlow, M.J.; Cubbon, R.M.; et al. Piezo1 Channels Sense Whole Body Physical Activity to Reset Cardiovascular Homeostasis and Enhance Performance. Nat. Commun. 2017, 8, 350. [Google Scholar] [CrossRef]
  159. Dora, K.A. Endothelial-Smooth Muscle Cell Interactions in the Regulation of Vascular Tone in Skeletal Muscle. Microcirculation 2016, 23, 626–630. [Google Scholar] [CrossRef]
  160. Genet, N.; Billaud, M.; Rossignol, R.; Dubois, M.; Gillibert-Duplantier, J.; Isakson, B.E.; Marthan, R.; Savineau, J.P.; Guibert, C. Signaling Pathways Linked to Serotonin-Induced Superoxide Anion Production: A Physiological Role for Mitochondria in Pulmonary Arteries. Front. Physiol. 2017, 8, 76. [Google Scholar] [CrossRef] [Green Version]
  161. Billaud, M.; Marthan, R.; Savineau, J.P.; Guibert, C. Vascular Smooth Muscle Modulates Endothelial Control of Vasoreactivity via Reactive Oxygen Species Production through Myoendothelial Communications. PLoS ONE 2009, 4, e6432. [Google Scholar] [CrossRef] [PubMed]
  162. McHugh, B.J.; Buttery, R.; Lad, Y.; Banks, S.; Haslett, C.; Sethi, T. Integrin Activation by Fam38A uses a Novel Mechanism of R-Ras Targeting to the Endoplasmic Reticulum. J. Cell Sci. 2010, 123, 51–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Solis, A.G.; Bielecki, P.; Steach, H.R.; Sharma, L.; Harman, C.C.D.; Yun, S.; de Zoete, M.R.; Warnock, J.N.; To, S.D.F.; York, A.G.; et al. Mechanosensation of Cyclical Force by PIEZO1 Is Essential for Innate Immunity. Nature 2019, 573, 69–74. [Google Scholar] [CrossRef]
  164. Sun, Y.; Leng, P.; Song, M.; Li, D.; Guo, P.; Xu, X.; Gao, H.; Li, Z.; Li, C.; Zhang, H. Piezo1 Activates the NLRP3 Inflammasome in Nucleus Pulposus Cell-Mediated by Ca2+/NF-kappaB Pathway. Int. Immunopharmacol. 2020, 85, 106681. [Google Scholar] [CrossRef]
  165. Albarran-Juarez, J.; Iring, A.; Wang, S.; Joseph, S.; Grimm, M.; Strilic, B.; Wettschureck, N.; Althoff, T.F.; Offermanns, S. Piezo1 and Gq/G11 Promote Endothelial Inflammation Depending on Flow Pattern and Integrin Activation. J. Exp. Med. 2018, 215, 2655–2672. [Google Scholar] [CrossRef] [Green Version]
  166. Fulton, D.J.R.; Li, X.; Bordan, Z.; Haigh, S.; Bentley, A.; Chen, F.; Barman, S.A. Reactive Oxygen and Nitrogen Species in the Development of Pulmonary Hypertension. Antioxidants 2017, 6, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Sun, Y.; Li, M.; Liu, G.; Zhang, X.; Zhi, L.; Zhao, J.; Wang, G. The Function of Piezo1 in Colon Cancer Metastasis and Its Potential Regulatory Mechanism. J. Cancer Res. Clin. Oncol. 2020, 146, 1139–1152. [Google Scholar] [CrossRef] [Green Version]
  168. Chen, X.; Wanggou, S.; Bodalia, A.; Zhu, M.; Dong, W.; Fan, J.J.; Yin, W.C.; Min, H.K.; Hu, M.; Draghici, D.; et al. A Feedforward Mechanism Mediated by Mechanosensitive Ion Channel PIEZO1 and Tissue Mechanics Promotes Glioma Aggression. Neuron 2018, 100, 799–815.e7. [Google Scholar] [CrossRef] [Green Version]
  169. Richter, K.; Kiefer, K.P.; Grzesik, B.A.; Clauss, W.G.; Fronius, M. Hydrostatic Pressure Activates ATP-Sensitive K+ Channels in Lung Epithelium by ATP Release through Pannexin and Connexin Hemichannels. FASEB J. 2014, 28, 45–55. [Google Scholar] [CrossRef]
  170. Xin, F.; Cheng, Y.; Ren, J.; Zhang, S.; Liu, P.; Zhao, H.; Huang, H.; Wang, W. The Extracellular Loop of the Auxiliary beta1-Subunit Is Involved in the Regulation of BKCa Channel Mechanosensitivity. Am. J. Physiol. Cell Physiol. 2018, 315, C485–C493. [Google Scholar] [CrossRef]
  171. Wiedmann, F.; Rinne, S.; Donner, B.; Decher, N.; Katus, H.A.; Schmidt, C. Mechanosensitive TREK-1 Two-Pore-Domain Potassium (K2P) Channels in the Cardiovascular System. Prog. Biophys. Mol. Biol. 2021, 159, 126–135. [Google Scholar] [CrossRef] [PubMed]
  172. Xu, Z.; Chen, Z.M.; Wu, X.; Zhang, L.; Cao, Y.; Zhou, P. Distinct Molecular Mechanisms Underlying Potassium Efflux for NLRP3 Inflammasome Activation. Front. Immunol. 2020, 11, 609441. [Google Scholar] [CrossRef] [PubMed]
  173. Mondejar-Parreno, G.; Cogolludo, A.; Perez-Vizcaino, F. Potassium (K+) Channels in the Pulmonary Vasculature: Implications in Pulmonary Hypertension Physiological, Pathophysiological and Pharmacological Regulation. Pharmacol. Ther. 2021, 225, 107835. [Google Scholar] [CrossRef]
  174. Tang, Q.Y.; Qi, Z.; Naruse, K.; Sokabe, M. Characterization of a Functionally Expressed Stretch-Activated BKca Channel Cloned from Chick Ventricular Myocytes. J. Membr. Biol. 2003, 196, 185–200. [Google Scholar] [CrossRef] [PubMed]
  175. Kirber, M.T.; Ordway, R.W.; Clapp, L.H.; Walsh, J.V., Jr.; Singer, J.J. Both Membrane Stretch and Fatty Acids Directly Activate Large Conductance Ca2+-Activated K+ Channels in Vascular Smooth Muscle Cells. FEBS Lett. 1992, 297, 24–28. [Google Scholar] [CrossRef] [Green Version]
  176. Ferraz, A.P.; Seara, F.A.C.; Baptista, E.F.; Barenco, T.S.; Sottani, T.B.B.; Souza, N.S.C.; Domingos, A.E.; Barbosa, R.A.Q.; Takiya, C.M.; Couto, M.T.; et al. BKCa Channel Activation Attenuates the Pathophysiological Progression of Monocrotaline-Induced Pulmonary Arterial Hypertension in Wistar Rats. Cardiovasc. Drugs Ther. 2020, 35, 719–732. [Google Scholar] [CrossRef]
  177. Huang, H.; Liang, L.; Liu, P.; Wei, H.; Sachs, F.; Niu, W.; Wang, W. Mechanical Effects on KATP Channel Gating in Rat Ventricular Myocytes. PLoS ONE 2013, 8, e63337. [Google Scholar] [CrossRef] [Green Version]
  178. Clapp, L.H.; Gurney, A.M. ATP-Sensitive K+ Channels Regulate Resting Potential of Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol. 1992, 262, H916–H920. [Google Scholar] [CrossRef]
  179. Cui, Y.; Tran, S.; Tinker, A.; Clapp, L.H. The Molecular Composition of K(ATP) Channels in Human Pulmonary Artery Smooth Muscle Cells and Their Modulation by Growth. Am. J. Respir. Cell Mol. Biol. 2002, 26, 135–143. [Google Scholar] [CrossRef]
  180. Koh, S.D.; Monaghan, K.; Sergeant, G.P.; Ro, S.; Walker, R.L.; Sanders, K.M.; Horowitz, B. TREK-1 Regulation by Nitric Oxide and cGMP-Dependent Protein Kinase. An Essential Role in Smooth Muscle Inhibitory Neurotransmission. J. Biol. Chem. 2001, 276, 44338–44346. [Google Scholar] [CrossRef] [Green Version]
  181. Gardener, M.J.; Johnson, I.T.; Burnham, M.P.; Edwards, G.; Heagerty, A.M.; Weston, A.H. Functional Evidence of a Role for Two-Pore Domain Potassium Channels in Rat Mesenteric and Pulmonary Arteries. Br. J. Pharmacol. 2004, 142, 192–202. [Google Scholar] [CrossRef] [Green Version]
  182. Xie, W.; Wang, H.; Ding, J.; Wang, H.; Hu, G. Anti-Proliferating Effect of Iptakalim, a Novel KATP Channel Opener, in Cultured Rabbit Pulmonary Arterial Smooth Muscle Cells. Eur. J. Pharmacol. 2005, 511, 81–87. [Google Scholar] [CrossRef] [PubMed]
  183. He, M.; Cui, T.; Cai, Q.; Wang, H.; Kong, H.; Xie, W. Iptakalim Ameliorates Hypoxia-Impaired Human Endothelial Colony-Forming Cells Proliferation, Migration, and Angiogenesis via Akt/eNOS Pathways. Pulm. Circ. 2019, 9, 2045894019875417. [Google Scholar] [CrossRef] [PubMed]
  184. Li, J.; Long, C.; Cui, W.; Wang, H. Iptakalim Ameliorates Monocrotaline-Induced Pulmonary Arterial Hypertension in Rats. J. Cardiovasc. Pharmacol. Ther. 2013, 18, 60–69. [Google Scholar] [CrossRef]
  185. Hayabuchi, Y. The Action of Smooth Muscle Cell Potassium Channels in the Pathology of Pulmonary Arterial Hypertension. Pediatr. Cardiol. 2017, 38, 1–14. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hemodynamic forces acting on the vessel wall. Section of an artery wall showing that the endothelial cells, forming the inner tunica, are longitudinally aligned, whereas smooth muscle cells, forming the median layer, are circumferentially aligned; the surrounding adventitia predominantly includes fibroblasts and matrix. Shear stress, frictional force generated by blood flow, is parallel to the vessel wall, whereas blood pressure is perpendicular to the vessel wall, causing circumferential and longitudinal stretching. Beside blood mechanical forces, composition of extracellular matrix, contributing to arterial stiffness, may itself modulate compliance and mechanotransduction in the vessel wall.
Figure 1. Hemodynamic forces acting on the vessel wall. Section of an artery wall showing that the endothelial cells, forming the inner tunica, are longitudinally aligned, whereas smooth muscle cells, forming the median layer, are circumferentially aligned; the surrounding adventitia predominantly includes fibroblasts and matrix. Shear stress, frictional force generated by blood flow, is parallel to the vessel wall, whereas blood pressure is perpendicular to the vessel wall, causing circumferential and longitudinal stretching. Beside blood mechanical forces, composition of extracellular matrix, contributing to arterial stiffness, may itself modulate compliance and mechanotransduction in the vessel wall.
Biomolecules 11 01389 g001
Figure 2. Activation mechanisms of SAC. Three general models are proposed: (a) In the “bilayer model”, the tension developed (red arrow) in the lipid bilayer itself is directly responsible for channel gating. (b) In the “tether model”, the force is transmitted to the channel via proteins located in the extracellular matrix, the cytoskeleton, or both. Tensions are conveyed by these accessory proteins to induce the channel opening. (c) In the “secondary signal model”, the channel activation depends on a distant mechanical-sensitive protein generating diffusible second messenger or channel phosphorylation.
Figure 2. Activation mechanisms of SAC. Three general models are proposed: (a) In the “bilayer model”, the tension developed (red arrow) in the lipid bilayer itself is directly responsible for channel gating. (b) In the “tether model”, the force is transmitted to the channel via proteins located in the extracellular matrix, the cytoskeleton, or both. Tensions are conveyed by these accessory proteins to induce the channel opening. (c) In the “secondary signal model”, the channel activation depends on a distant mechanical-sensitive protein generating diffusible second messenger or channel phosphorylation.
Biomolecules 11 01389 g002
Figure 3. Experimental strategies to investigate SAC in cells. At the cellular level, several strategies can be used to activate SAC. The most commonly used are based on membrane deformation: (a) applying positive or negative pressure to the back end of the patch pipette, (b) poking of the cell membrane by a piezo-driven glass pipette, (c) modifying the perfusion flow or the viscosity of the solution, (d) using osmotic challenges: hypotonicity induces cell swelling, whilst hypertonicity evokes cell-shrinkage, (e) elongating thin elastic silicone membrane where cells are seeded, (f) applying magnetic field to specific ligands coated with magnetic particles on the cells, (g) seeding cells on elastomeric pillars to apply force to specific parts of the cells, and (h) using crenators and cup formers (amphipathic compounds) to induce crenation or cup shapes. (i) Another alternative consists in culturing cells in matrices of different stiffness, to evaluate the impact of the environment matrix and more especially its stiffness.
Figure 3. Experimental strategies to investigate SAC in cells. At the cellular level, several strategies can be used to activate SAC. The most commonly used are based on membrane deformation: (a) applying positive or negative pressure to the back end of the patch pipette, (b) poking of the cell membrane by a piezo-driven glass pipette, (c) modifying the perfusion flow or the viscosity of the solution, (d) using osmotic challenges: hypotonicity induces cell swelling, whilst hypertonicity evokes cell-shrinkage, (e) elongating thin elastic silicone membrane where cells are seeded, (f) applying magnetic field to specific ligands coated with magnetic particles on the cells, (g) seeding cells on elastomeric pillars to apply force to specific parts of the cells, and (h) using crenators and cup formers (amphipathic compounds) to induce crenation or cup shapes. (i) Another alternative consists in culturing cells in matrices of different stiffness, to evaluate the impact of the environment matrix and more especially its stiffness.
Biomolecules 11 01389 g003
Figure 4. Experimental strategies to investigate SAC in vessels. At the tissue level, the effects of stretch can also be studied in whole vessels using (a) arteriography: the microvessel is cannulated at both ends with glass micropipettes and placed in a microvascular flow system chamber, allowing intraluminal pressure increase via modulation of inlet and outlet pressures (Pin and Pout, respectively); or (b) myography: one end of the segment is anchored to a stationary support and the other end is connected to a force-displacement transducer to monitor the vessel contraction under resting tension corresponding to an adapted transmural pressure.
Figure 4. Experimental strategies to investigate SAC in vessels. At the tissue level, the effects of stretch can also be studied in whole vessels using (a) arteriography: the microvessel is cannulated at both ends with glass micropipettes and placed in a microvascular flow system chamber, allowing intraluminal pressure increase via modulation of inlet and outlet pressures (Pin and Pout, respectively); or (b) myography: one end of the segment is anchored to a stationary support and the other end is connected to a force-displacement transducer to monitor the vessel contraction under resting tension corresponding to an adapted transmural pressure.
Biomolecules 11 01389 g004
Figure 5. Schematic view illustrating the multifunctional contribution of SAC in the pathogenesis of PH. Red arrows indicate PH-induced modifications of cellular processes in pulmonary arterial vascular cells.
Figure 5. Schematic view illustrating the multifunctional contribution of SAC in the pathogenesis of PH. Red arrows indicate PH-induced modifications of cellular processes in pulmonary arterial vascular cells.
Biomolecules 11 01389 g005
Table 1. Biophysical properties of TRP channels in pulmonary arterial vascular cells.
Table 1. Biophysical properties of TRP channels in pulmonary arterial vascular cells.
ChannelCell Type
(Species)
Conductance (pS)PermeabilityActivatorInhibitorPA Phenotype in KO MiceReferences
TRPA1-9–16
53.1–62.8
-osmolarity
mustard oil
HC030031not described in PA[35]
TRPC1human, mouse, and rat PA (PAEC and PASMC)16PCa/PNa < 10stretch
store depletion
2-APB
Gd3+/La3+
reduced hyperreactivity, remodeling, and vasomotor tone[36,37,38,39]
TRPC3rat PA
human PASMC
66PCa/PNa = 1.6stretch
store depletion
DAG and analogs
2-APB
Gd3+/La3+
not described in PA[36,37,38,39]
TRPC4human, mouse, and rat PA (PAEC and PASMC)17.5–41PCa/PNa = 1.1–7store depletion, arachidonic acid
calmidazolium
2-APB
SKF96365
Gd3+/La3+
niflumic acid DIDS
reduced vascular permeability
(and remodeling in KO rats)
[40,41,42,43,44,45,46,47]
TRPC5human, mouse, and rat PA
human PASMC
64PCa/PNa = 9stretch
hypotonocity
roziglitazone
2-APB
La3+
not described in PA[36,38,39,48,49,50]
TRPC6human, mouse, and rat PA (PAEC and PASMC)28–37PCa/PNa = 4–5stretch
hypotonicity
store depletion
DAG and analogs
2-APB
SKF96365
Cd2+, La3+, Gd3+
reduced hyperreactivity, remodeling
and vasomotor tone
[36,37,38,39]
TRPM3rat PA65–133PCa/PNa = 1.5–2hypotonicity not described in PA[51,52]
TRPM4rat PA24–25Na+//K+ > Cs+ > Li+stretch not described in PA[52,53,54,55]
TRPM7rat PA
human and rat PASMC
105
22–37
Zn2+ ≈ Ni2+ >> Ba2+ > Co2+ > Mg2+ ≥ Mn2+ ≥ Sr2+ ≥ Cd2+ ≥ Ca2+
PNa/PCa = 3
stretch
osmolarity
not viable
→ conditional deletion: reduced remodeling and vasomotor tone
[52,56,57,58,59]
TRPP1-135–175-PKD1
stretch
-not described in PA[36,60]
TRPP2-177PCa/PNa = 1–5shear stress
hypotonicity
triptolide
Gd3+/La3+not described in PA
(conditional in smooth muscle cells)
[36,37,61,62,63]
TRPV1human and rat PA
(PAEC and PASMC)
35–80
(unitary)
143–144 (whole cell)
PCa/PNa = 10 (capsaicin-activated)hypotonicity
capsaicin
2-APB
cannabidiol
resiniferatoxin
capsazepine
AMG9810
A784168
5’-Iodoresiniferatoxin
not described in PA[36,52,64]
TRPV2human and rat PASMC-PCa/PNa = 1–3hypotonicity
stretch
2-APB
tranilastnot described in PA[36,52]
TRPV4human, mouse, and rat PA (PAAF PAEC, PASMC)30–90
(unitary)
452 ± 63 (whole cell)
PCa/PNa = 6–10hypotonicity
shear stress
4-αPDD
GSK1016790A
HC607047
RN-1734
GSK2193874
reduced hyperreactivity, remodeling
and vasomotor tone
[23,36,52,65,66]
Table 2. Biophysical properties of Piezo channels in pulmonary arterial vascular cells.
Table 2. Biophysical properties of Piezo channels in pulmonary arterial vascular cells.
ChannelCell Type
(Species)
Conductance (pS)PermeabilityActivatorInhibitorPA Phenotype in KO MiceReferences
Piezo1human, mouse, and rat PA
human and mouse PAEC
human PASMC
22–30Ca2+ > Na+, K+, Mg2+
PCa/PCs = 2; PNa/PCs = 1.1; PK/PCs = 1.1; PMg/PCs = 0.5
negative and positive pressures, shear stress, ultrasound waves, Yoda1, Jedi1/2GsMTx-4
ruthenium red
Gd3+
dooku1
tubeimoside I
not viable
→ endothelium specific conditional deletion: no effect
[26,132,134,135,137,138,139,140,141,142,144,145,146,147,148,149,150]
Piezo2human PA27–28-negative and positive pressuresGsMTx-4
ruthenium red
Gd3+
not described in PA[132,136,138,143,147]
Table 3. Biophysical properties of mechanosensitive K+ channels in pulmonary arterial vascular cells.
Table 3. Biophysical properties of mechanosensitive K+ channels in pulmonary arterial vascular cells.
ChannelCell Type
(Species)
Conductance (pS)PermeabilityActivatorInhibitorPA Phenotype in KO MiceReferences
BKCarabbit and rat PASMC273K+negative pressure, calcium-not described in PA[170,174,175,176]
KATPhuman and rabbit PASMC42–55 (pressure), 28 (Levcromakalim)K+negative and positive pressures, levcromakalim, iptakalimintracellular ATP
glibenclamide
not described in PA[177,178,179]
TREK-1mouse PA
rat PASMC
90K+negative pressure-not described in PA[180,181]
TREK-2mouse and rat PA-K+stretch-not described in PA[180,181]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Barbeau, S.; Gilbert, G.; Cardouat, G.; Baudrimont, I.; Freund-Michel, V.; Guibert, C.; Marthan, R.; Vacher, P.; Quignard, J.-F.; Ducret, T. Mechanosensitivity in Pulmonary Circulation: Pathophysiological Relevance of Stretch-Activated Channels in Pulmonary Hypertension. Biomolecules 2021, 11, 1389. https://doi.org/10.3390/biom11091389

AMA Style

Barbeau S, Gilbert G, Cardouat G, Baudrimont I, Freund-Michel V, Guibert C, Marthan R, Vacher P, Quignard J-F, Ducret T. Mechanosensitivity in Pulmonary Circulation: Pathophysiological Relevance of Stretch-Activated Channels in Pulmonary Hypertension. Biomolecules. 2021; 11(9):1389. https://doi.org/10.3390/biom11091389

Chicago/Turabian Style

Barbeau, Solène, Guillaume Gilbert, Guillaume Cardouat, Isabelle Baudrimont, Véronique Freund-Michel, Christelle Guibert, Roger Marthan, Pierre Vacher, Jean-François Quignard, and Thomas Ducret. 2021. "Mechanosensitivity in Pulmonary Circulation: Pathophysiological Relevance of Stretch-Activated Channels in Pulmonary Hypertension" Biomolecules 11, no. 9: 1389. https://doi.org/10.3390/biom11091389

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop