Next Article in Journal / Special Issue
The Non-Fibrillar Side of Fibrosis: Contribution of the Basement Membrane, Proteoglycans, and Glycoproteins to Myocardial Fibrosis
Previous Article in Journal
Arterial Structural and Functional Characteristics at End of Early Childhood and Beginning of Adulthood: Impact of Body Size Gain during Early, Intermediate, Late and Global Growth
Previous Article in Special Issue
Cardiac Fibroblasts and the Extracellular Matrix in Regenerative and Nonregenerative Hearts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ca2+ Signaling in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases

1
Calhoun Cardiology Center, Department of Cell Biology, University of Connecticut Health Center, Farmington, CT 06030, USA
2
Department of Pharmacology, Rutgers, Robert Wood Johnson Medical School, Piscataway, NJ 08854, USA
*
Authors to whom correspondence should be addressed.
J. Cardiovasc. Dev. Dis. 2019, 6(4), 34; https://doi.org/10.3390/jcdd6040034
Submission received: 11 August 2019 / Revised: 16 September 2019 / Accepted: 18 September 2019 / Published: 23 September 2019
(This article belongs to the Special Issue Cardiac Fibroblasts and Fibrosis)

Abstract

:
Cardiac fibrosis is the excessive deposition of extracellular matrix proteins by cardiac fibroblasts and myofibroblasts, and is a hallmark feature of most heart diseases, including arrhythmia, hypertrophy, and heart failure. This maladaptive process occurs in response to a variety of stimuli, including myocardial injury, inflammation, and mechanical overload. There are multiple signaling pathways and various cell types that influence the fibrogenesis cascade. Fibroblasts and myofibroblasts are central effectors. Although it is clear that Ca2+ signaling plays a vital role in this pathological process, what contributes to Ca2+ signaling in fibroblasts and myofibroblasts is still not wholly understood, chiefly because of the large and diverse number of receptors, transporters, and ion channels that influence intracellular Ca2+ signaling. Intracellular Ca2+ signals are generated by Ca2+ release from intracellular Ca2+ stores and by Ca2+ entry through a multitude of Ca2+-permeable ion channels in the plasma membrane. Over the past decade, the transient receptor potential (TRP) channels have emerged as one of the most important families of ion channels mediating Ca2+ signaling in cardiac fibroblasts. TRP channels are a superfamily of non-voltage-gated, Ca2+-permeable non-selective cation channels. Their ability to respond to various stimulating cues makes TRP channels effective sensors of the many different pathophysiological events that stimulate cardiac fibrogenesis. This review focuses on the mechanisms of Ca2+ signaling in fibroblast differentiation and fibrosis-associated heart diseases and will highlight recent advances in the understanding of the roles that TRP and other Ca2+-permeable channels play in cardiac fibrosis.

1. Introduction

Cardiac fibrosis is involved in pathological remodeling of the heart, causing abnormalities in cardiac conduction, stiffness of the ventricular walls, reduced contractility, and impaired overall heart performance [1]. Thus, cardiac fibrosis is a detrimental factor in various types of heart diseases [2,3,4], including arrhythmia [5,6,7,8,9], hypertrophy [10], and heart failure [10,11,12,13]. There is substantial experimental evidence demonstrating that fibrosis slows down action potential propagation, initiates reentry, and promotes ectopic automaticity, thereby contributing to arrhythmogenesis [5,9,14]. Fibrosis also accelerates the progression of heart failure [15,16] resulting from nearly all etiologies of heart diseases, such as ischemic cardiomyopathy [17], dilated cardiomyopathy [18,19,20,21], hypertensive heart diseases [22,23], and inflammatory heart diseases [24]. Therefore, mitigating cardiac fibrosis by targeting the fibrogenesis cascade offers numerous promising therapeutic strategies for fibrosis-associated heart diseases [2,25].
Fibrosis is the excessive accumulation of extracellular matrix (ECM) proteins synthesized by fibroblasts and myofibroblasts. The differentiation of fibroblasts into myofibroblasts is a pivotal step in the fibrogenesis cascade, as myofibroblasts are the predominant cell type that synthesizes and secretes ECM proteins [26]. Moreover, myofibroblasts produce growth factors, cytokines, and metalloproteinases, and are particularly responsive to proinflammatory cytokines, including tumor necrosis factor α (TNFα); interleukin-1 (IL-1), IL-6, and TGF-β; vasoactive peptide AngII, ET-1, ANP and BNP [27]; and hormones, such as noradrenaline. Thus, the differentiation of quiescent fibroblasts to active matrix-producing myofibroblasts is a key step in disease progression. A variety of pathological factors, such as oxidative stress, myocardial injury, unbalanced hormone levels, mechanical overload, and inflammatory stimuli can promote fibroblasts to differentiate into myofibroblasts. Increasing the differentiation of fibroblasts is, therefore, essential for initiating and perpetuating the fibrogenesis cascade [28,29,30,31,32,33,34] and the formation of fibrosis, which is involved in various pathological cardiac remodeling processes [9,35] (Figure 1). Numerous signaling pathways are involved in the activation of cardiac fibrogenesis [36], among which intracellular Ca2+ has been found to play a particularly critical role [37,38,39]. The Ca2+ signaling mechanisms in cardiac fibroblasts, however, are not fully understood. In recent years, there has been an increasing interest in elucidating the Ca2+ signaling mechanisms in cardiac fibroblasts and myofibroblasts. Similar to other cell types, Ca2+ signals in cardiac fibroblasts are generated by both Ca2+ release and Ca2+ entry. However, as non-excitable cells in the heart, cardiac fibroblasts are in a unique, complex environment, surrounded by multiple stimulating cues for fibrotic response. Recently, TRP channels have been found to be essential to fibroblast Ca2+ signaling. As Ca2+-permeable non-selective cation channels that respond to a variety of stimuli, TRP channels can not only directly initiate cellular Ca2+ signaling, but can also cause membrane depolarization to indirectly influence Ca2+ influx via other Ca2+-permeable channels. Moreover, although not gated by voltage, TRP channel activity can be influenced by membrane potentials. In addition, various fibrotic stimuli are known to activate Gq-coupled receptors in fibroblasts to induce Ca2+ influx through store-operated ion channels and TRP channels. This short review summarizes recent advances in the understanding of the different components contributing to intracellular Ca2+ signals, including Ca2+ release and Ca2+ entry, and their roles in cardiac fibrosis and fibrosis-associated heart diseases (Figure 2).

2. Ion Channels Controlling the Membrane Potential in Cardiac Fibroblasts

The membrane potential is a major force that controls Ca2+ entry in all different types of cells. Cardiac fibroblasts have a more depolarized resting membrane potential in comparison to cardiac myocytes. Measured by standard microelectrode techniques in multicellular tissues, the resting membrane potential of atrial fibroblasts was reported to be between –31 and –16 mV [40,41,42]. A similar resting membrane potential (about –37mV to –40 mV) was also obtained by patch-clamp recordings in isolated rat atrial fibroblasts [43,44]. The resting membrane potential was found to be more hyperpolarized in fibroblasts derived from canine heart failure models [44]. For example, it was increased from –43 mV in control fibroblasts to –56 mV in heart failure fibroblasts [44], pointing to the altered regulation of ion channels under diseased conditions [44]. Although fibroblasts are non-excitable cells, early studies have shown evidence that functional voltage-gated ion channels are expressed in cardiac fibroblasts [45]. Indeed, a growing number of studies have suggested the importance of both voltage- and non-voltage-gated ion channels in fibroblast function.

2.1. K+ Channels

The expression of many different ion channel genes in cardiac fibroblasts has been detected at the RNA level by PCR and by patch-clamp current recordings in various studies (Table 1) [36,46]. Several voltage-gated ion channels and non-voltage-gated ion channels are present in cardiac fibroblasts (Table 1). For voltage-gated K+ channels, the transient-outward K+-current (Ito), and delayed rectifier K+-currents have been demonstrated in rat [47,48], canine [49,50], and human [51] fibroblasts. These potassium channels are likely to be encoded by the α-subunits of voltage-gated K+ channel proteins, such as Kv1.2, Kv1.4, Kv1.5, and Kv2.1. In addition to the voltage-gated K+ channels, inward rectifier K+ (Kir or IK1) currents have been recorded in rat [52], canine [53], and human [51] fibroblasts. IK1 activity has been reported to influence myofibroblast proliferation and contraction functions during in vitro experiments [52]. Upregulated IK1 has also been reported to play a role in atrial remodeling during atrial fibrillation in a canine model of congestive heart failure [53]. Another significant K+ conductance in fibroblasts is the ATP-activated potassium channel, KATP [51,52,54]. Activation of KATP in mouse cardiac fibroblasts increases cell proliferation, reduces IL-6 secretion [54], and inhibits fibroblast differentiation induced by ischemic injury in vitro [55]. Functional KATP expression in fibroblasts in vivo can be induced by myocardial infarction in the scar and border zone, which may influence the electrophysiological properties of myocytes [56]. Several types of Ca2+-activated K+ currents, including the big conductance (BKCa) [51,57,58] and intermediate conductance (KCa3.1) [59], have been demonstrated in human and rat fibroblasts; these currents play a role in fibroblast proliferation [58], fibroblast-myocyte coupling [57], and the response to stretch [60].

2.2. Na+ Channels

Voltage-gated Na+ channels have been reported to be expressed in cardiac fibroblasts [51,58,61,62,63]. Both tetrodotoxin-sensitive and TTX-resistance Na+ currents (INa) were recorded in cultured human ventricular fibroblasts [51]. In cultured human atrial fibroblasts, TTX-resistant INa was only identified in differentiated myofibroblasts, but not in freshly isolated fibroblasts [61]. The TTX-resistant INa exhibits similar properties to that of INa in cardiac myocytes, and is likely encoded by NaV1.5’s α-subunit and β1-subunit, as the expression levels of these two subunits were markedly higher in myofibroblasts than in fibroblasts [61]. Interestingly, although larger INa was found to present in a greater number of fibroblasts from atrial fibrillation patients than those from normal rhythm patients [63], INa did not appear to influence fibroblast proliferation, differentiation, or migration [61]. Moreover, as different α-subunits of Na+ channels are expressed in fibroblasts, including Nav1.2, NaV1.3, NaV1.5, Nav1.6, Nav1.7, and NaV1.9 [51,61,62], INa may be generated by a combination of different isoforms of the α subunits. Further investigation is required to understand the role of INa in fibroblasts’ and myofibroblasts’ functions and fibroblast-myocyte coupling.

2.3. Cl and H+ Channels

Swelling-induced Cl currents [51,58] are present in human cardiac fibroblasts and are involved in fibroblast proliferation. Additionally, functional voltage-gated H+ currents are present in cardiac fibroblasts [66]. It was proposed that H+ currents in cardiac fibroblasts are involved in the regulation of intracellular pH and membrane potential under physiological conditions, as well as in the response to pathological conditions, such as ischemia, although their purpose has yet to be determined [66,74].
In summary, with growing interest in understanding the role of fibroblasts in cardiac physiology and pathology, the knowledge of cardiac fibroblasts’ electrophysiological properties will help elucidate how these properties influence the functions of cardiac fibroblasts, including the proliferation, differentiation, and secretion of ECM proteins, as well as the couplings of fibroblasts and myocytes, as reviewed recently [75]. The electrophysiological properties of fibroblasts can also directly influence Ca2+ signaling in fibroblasts, which plays a critical role in the cardiac fibrogenesis cascade.

3. Ca2+ Signaling Mechanisms in Cardiac Fibroblasts

As a ubiquitous second messenger, Ca2+ is essential for nearly all cellular functions, including cell signaling, gene expression, cell proliferation, differentiation, migration, growth, and death [76,77]. Intracellular Ca2+ signals are generated by Ca2+ entry through Ca2+-permeable channels in the plasma membrane and Ca2+ release from intracellular Ca2+ stores. Intracellular Ca2+ levels are also finely controlled by plasma membrane ATPases (PMCAs), Na+/Ca2+ exchangers (NCX), and the sarcoplasmic reticulum Ca2+-ATPase (SERCA) [78]. Unlike cardiac myocytes, whose electrical properties and Ca2+ signaling mechanisms are well understood, the knowledge of Ca2+ signaling in cardiac fibroblasts is limited. In recent years, however, there have been extensive studies investigating the role of Ca2+ signaling in fibroblast physiology, and in fibrosis-associated heart diseases.

3.1. Ca2+ Release Mechanisms in Cardiac Fibroblasts

There are several Ca2+ release mechanisms in different types of cells [78,79]. Ca2+ release from intracellular stores by ryanodine receptor (RyR) and IP3 receptor (IP3R) activation represent two major Ca2+ release mechanisms. There are also RyR-like Ca2+ release channels activated by cyclic ADP-ribose (cADPR) [78,79], as well as a distinct Ca2+-release pathway activated by nicotinic acid adenine dinucleotide phosphate (NAADP) [78,79]. NAADP is a potent intracellular Ca2+ release messenger originally identified in sea urchin lysates [80]. The target receptors of NAADP have been proposed to be RyR1 [81], and the two-pore channels (TPC) TPC1 and TPC2 [82,83,84,85]. However, whether TPC1 and TPC2 are the target receptors for NAADP is still controversial and requires further investigation [86].
Among the Ca2+ release pathways in the cardiac fibroblasts, no apparent role for RyRs has been demonstrated thus far. Although RyR blockers and activators have been reported to influence mechanically-induced potentials (MIP) recorded in fibroblasts in rat right atria [87], RyR expression, including that of RyR1, RyR2, and RyR3, cannot be detected by PCR in cultured human cardiac fibroblasts [46] or in neonatal rat cardiac fibroblasts [88]. Moreover, Ca2+ oscillations observed in cultured human cardiac fibroblasts are not influenced by ryanodine or caffeine [46]. Nonetheless, RyR2 expressed in myocytes has been shown to be involved in regulating TGF-β levels, thereby influencing fibrosis formation [88]. These results indicate that RyR may not have a direct role in fibroblasts under physiological and pathological conditions [88].
Similar to RyRs, which are known to be essential for the function of cardiac myocytes, the role of NAADP- and cADPR-induced Ca2+ signaling in cardiac myocytes has been demonstrated in various studies [89,90,91,92]. NAADP- and cADPR-mediated Ca2+ signaling contributes to the development of maladaptive cardiac hypertrophy induced by β-adrenergic stimulation [90]. However, although NAADP- and cADPR-mediated Ca2+ signaling has been shown to be crucial in various cell types, the role of these receptors in cardiac fibroblasts, if any, has not been determined.
Therefore, among different intracellular Ca2+ release pathways, IP3R activation-mediated Ca2+ signaling appears to be the major Ca2+ release pathway in cardiac fibroblasts. Chen and colleagues thoroughly investigated the expression of different components of Ca2+ release mechanisms in human cardiac fibroblasts, and found that all three types of IP3Rs are expressed in cultured human cardiac fibroblasts [46]. Moreover, Ca2+ oscillations can be completely blocked by the PLC blocker U73122 and the IP3R inhibitor 2-aminoethoxydiphenyl borate (2-APB). The IP3R agonist thimerosal can enhance Ca2+ oscillations [46]. Thus, it appears that IP3R activation-mediated Ca2+ release is the major Ca2+ release pathway in cardiac fibroblasts.

3.2. Signaling Pathways Mediating Ca2+ Release in Cardiac Fibroblasts

Many G-protein-coupled receptors (GPCRs) are expressed in cardiac fibroblasts [93]. AngII receptor AT1’s activation-induced intracellular Ca2+ increase has been proposed to be the signal transduction pathway responsible for AngII-mediated collagen synthesis [94]. Another Gq-coupled receptor activation by agonists has also been demonstrated to induce intracellular IP3 production in rat cardiac fibroblasts [95]. In addition to AngII, bradykinin (BK), ATP and UTP are also efficacious at inducing IP3 production, whereas endothelin-1 (ET1) and carbachol (100 μM) produce minimal or no change in IP3 production [95]. Accordingly, AngII (1 μM), BK (1 μM), UTP (30 μM), and ATP (30 μM) have been reported to induce substantial Ca2+, transiently, in rat cardiac fibroblasts, indicating that functional Gq-coupled receptors for AngII, BK, ATP, and UTP are present in cardiac fibroblasts [95].
The critical role of the AngII signaling pathway through fibroblast AT1 receptors in cardiac fibrosis has been well demonstrated [94,96,97,98]. Likewise, bradykinin receptors, Gi-coupled BR1, and Gq-coupled BR2 have been shown to be expressed in cardiac myofibroblasts [99,100]. In addition, purinergic signaling has also been shown to play an important role in cardiac fibrosis [101,102]. For the P1 receptors (adenosine receptors), RT-PCR has shown that mRNAs for all four P1 receptor subtypes, A1AR, A2AAR, A2BAR, and A3AR, are expressed in rat cardiac fibroblasts, with A2BARs being the most abundant [102]. Activation of A2BAR, which is coupled to Gq and Gs, inhibits collagen synthesis, connective tissue growth factor (CTGF) expression, and fibroblast proliferation [103]. Moreover, overexpression of A2B adenosine receptors results in a decrease in basal levels of collagen and protein synthesis, whereas silencing A2BARs results in an increase in protein and collagen synthesis [102]. Activation of the P2Y receptors can also modulate fibroblast function. For example, agonists of P2Y2 receptors activate human cardiac fibroblast proliferation and migration [104]. Among the eight purinergic P2Y receptors (P2YRs) [101,105], five Gq-coupled receptors (P2Y1, P2Y2, P2Y4, P2Y6, and P2Y11) can be detected by RT-PCR, immunostaining, and a functional assay that measures inositol phosphate (IP) production in rat myofibroblasts [106].
Recently, it has been demonstrated that the calcium-sensing receptor (CaSR) is expressed in cardiac fibroblasts [107]. CaSR acts as a G-protein-coupled receptor that can induce an increase of intracellular Ca2+ via Gq-PLC activation. CaSR is expressed in various cell types, including cardiac myocytes, smooth muscle cells, neurons, and vascular endothelial cells. In cardiac fibroblasts, activation of CaSR increases intracellular Ca2+ concentration, promotes fibroblast proliferation and migration, and induces the synthesis of extracellular matrix proteins [107]. Moreover, the inhibition of CaSR reduced fibrosis in an isoproterenol-induced rat hypertrophy model [107]. Therefore, CaSR appears to be important for mediating Ca2+ release in cardiac fibroblasts (Figure 2).

3.3. Ca2+ Entry Mechanisms in Cardiac Fibroblasts

Upon Gq-coupled receptor activation, Ca2+ release from intracellular stores is followed by Ca2+ entry through ion channels in the plasma membrane, which occurs through several different mechanisms. One pathway is mediated by store-operated Ca2+ entry (SOCE) through Ca2+ release-activated Ca2+ channels (CRAC). The second pathway occurs through non-selective Ca2+ entry after receptor activation, which is called receptor-operated Ca2+ entry (ROCE). ROCE is chiefly mediated by non-selective TRP channels [108], as discussed below. In addition to these two types of Ca2+ entry, purinergic P2X receptors (P2X1-7) can also directly mediate Ca2+ influx into cells. Although voltage-gated Ca2+ channel expression can be detected by PCR [46], functional currents have not been reported in cardiac fibroblasts.

3.3.1. Voltage-gated Ca2+ Channels

Although it has been demonstrated that both voltage-gated Na+ channels and K+ channels are functionally expressed in cardiac fibroblasts, voltage-gated Ca2+ currents have not been identified. The expression of CaV1.2 has been detected at the RNA expression level in human cardiac fibroblasts [46]. However, the functional voltage-gated Ca2+ channels cannot be obtained by patch-clamp [63]. Moreover, the Ca2+ channel blocker nifedipine and activator Bay K8644 failed to affect Ca2+ oscillations in cardiac fibroblasts, suggesting that L-type Ca2+ channels are not involved in Ca2+ signaling in this cell type [46]. Interestingly, a potential role for the voltage-gated Ca2+ channels has been suggested in lung fibrosis, since nifedipine inhibits Ca2+ oscillations and attenuates bleomycin-induced lung fibrosis [109]. These findings suggest that fibroblasts in different organs and tissues may have varied molecular components and exhibit distinct properties.

3.3.2. Ca2+ Entry Mediated by P2X Receptors

Another type of Ca2+-permeable channel identified in cardiac fibroblasts is the family of ATP-activated Ca2+-permeable non-selective cation channels, P2Xs, or P2X receptors (P2XR: P2X1-7). P2X receptors are widely distributed in excitable and nonexcitable cells [110]. The non-selective cation permeation of P2XRs not only brings Ca2+ into the cells, but also causes depolarization. Among the seven members of the P2XR family, P2X7 is present in human primary skin fibroblasts [111]. P2X7 has been shown to be involved in tissue fibrosis [112,113]. Since P2X7 is one of the major mediators of inflammasome activation, the mechanisms by which P2X7 activation causes fibrosis are unclear. In cardiac fibroblasts, P2X4 and P2X7 mRNAs have been detected in cultured human ventricular fibroblasts [46]. Furthermore, cardiac myocyte-specific overexpression of the P2X4 receptor in mice decreases fibrosis in a hypertensive heart failure mouse model [114,115], yet it is unknown whether this is due to direct cardiac myocyte protection or the interaction between cardiac myocytes and fibroblasts [116]. The function of P2X4 and P2X7 in mediating Ca2+ entry in cardiac fibroblasts and their role in cardiac fibrosis needs further investigation.

3.3.3. Ca2+ Entry through CRAC (Orai/STIM) Channels in Cardiac Fibroblasts

The Ca2+ release-activated Ca2+ channel (CRAC) is essential for cellular functions in a variety of cells [117,118,119]. The molecular components of CRAC channels consist of the pore-forming subunit Orai (Orai1, Orai2, and Orai3), and the Ca2+ release-sensing subunit STIM (STIM1 and STIM2). The role of Orai/STIM in cardiac myocytes and cardiac function has been extensively studied [120,121,122,123]. However, controversial results have been reported regarding the contribution of Orai/STIM to cardiac physiology and pathology [120,124]. In fibroblasts, although the Orai and STIM subunits have been detected by qPCR or RT-PCR in mouse and human cardiac fibroblasts [37,46], CRAC currents have not been recorded by patch-clamp [37]. Recently, the functional Orai1/STIM1 channels have been demonstrated by Ca2+ imaging measurements in human ventricular fibroblasts [67,68,125]. Ca2+ influx, and the expression of Orai1 but not STIM1, have been shown to be much larger in ventricular fibroblasts derived from failing hearts compared to those of non-failing hearts [68]. In aged ventricular fibroblasts, while store-operated Ca2+ release and entry were shown to increase, the expressions of Orai1 and STIM1 remained unchanged [67]. These results suggest that Orai/STIM channels play a role in fibroblast Ca2+ signaling, and may therefore contribute to cardiac remodeling under pathological conditions. Nonetheless, further investigations using the fibroblast-specific deletion of Orai/STIM proteins may help to clarify the physiological and pathological roles of Orai/STIM in the heart.

3.3.4. Ca2+ Entry Mediated by TRP Channels in Cardiac Fibroblasts and Fibrotic Heart Disease

Overview of TRP Channels

Transient receptor potential (TRP) channels belong to a superfamily of non-voltage-gated Ca2+-permeable ion channels that can be divided into six subfamilies, including the TRPC (canonical), TRPV (vanilloid), TRPM (melastatin), TRPA (ankyrin), TRPP (polycystin), and TRPML (mucolipin) subfamilies [126,127]. There are different members in each subfamily; for example, the TRPC, TRPV, and TRPM subfamilies contain seven (TRPC1-7), six (TRPV1-6), and eight (TRPM1-8) members, respectively [126,127,128]. The TRPA subfamily has one channel protein [129,130,131]; the TRPP and TRPML subfamilies contain three channel proteins each [127,132]. The TRPP (polycystin) and TRPML (mucolipin) subfamilies are intracellular ion-channels [133], whereas the other TRP subfamilies appear to function at the plasma membrane. Most TRP channels are Ca2+-permeable non-selective cation channels, with the exceptions of TRPM4 and TRPM5, which are monovalent cation-selective channels [126,127], and TRPV5 and TRPV6, which are highly Ca2+-selective (PCa/PNa>100) [126,127,134,135]. TRPM6 and TRPM7 are unique channel proteins because they have a kinase domain at their C-terminals. Both TRPM6 and TRPM7 are permeable to Ca2+, Mg2+, Zn2+, and other trace metals [72,136,137,138].

TRP Channels are Multifunctional Cellular Sensors

TRP channels are not gated by voltage, but rather, these channels are responsive to a wide range of stimuli in a polymodal activation manner, including thermal, mechanical, oxidative, chemical, and nociceptive stresses, and local autocrine or paracrine environmental cues [127,139]. TRPV1–TRPV4 and TRPM3 are activated by high temperatures, whereas TRPM8, TRPA1, and TRPC5 are turned on at low temperatures [140,141,142]. Many of the TRPC channels, including TRPC2 [143], TRPC3 [144], TRPC6 [144], and TRPC7 [145], are activated by Gq-linked receptor activation; meanwhile, some of the TRPCs can be directly activated by diacylglycerol (DAG) [146]. Different TRPCs can form heterotetrameric channels, such as the TRPC1/TRPC4/TRPC5 and TRPC3/TRPC6/TRPC7 functional channels [147]. The Mg2+-permeable TRPM6 and TRPM7 are activated by lowering the intracellular free Mg2+ concentration [136,137,148,149,150], and these channels are likely to be constitutively active to a small degree under physiological conditions. Changes in intracellular Ca2+ ([Ca2+]i) influence the activation of many TRP channels. A rise of intracellular Ca2+ ([Ca2+]i) activates several TRP channels, including TRPM4 [151], TRPM5 [152,153], TRPM2 [154], and TRPA1 [155]; in contrast, a decrease of intracellular Ca2+ ([Ca2+]i) can activate TRPV5 and TRPV6. TRPM2 can also be activated by multiple stimuli, including [Ca2+]i, ADP-ribose (ADPR), NAD+, and oxidative stress [156,157,158]. The polymodal activation feature of TRP channels makes them extremely powerful effectors for integrating and responding to a range of physiological and pathological stimuli, which in turn can trigger various pathogenesis cascades.

The Functional Expressions of the TRP Channels in Cardiac Fibroblasts

Many TRP channels have been detected at the RNA level by qPCR or RT-PCR in the cardiac fibroblasts of various species [37,46,159]. TRPC1, TRPC2, TRPC3, TRPC5, TRPC6, and TRPC7 are present in isolated rat fibroblasts [69,160]; TRPC1, TRPC4, and TRPC6 are expressed in cultured human cardiac fibroblasts [46]; and TRPC1, TRPC6, TRPV2, TRPV4, and TRPM7 can be detected in freshly isolated human atrial fibroblasts by RT-PCR [37]. In mouse cardiac fibroblasts, TRPC1, TRPC3, TRPC4, TRPC6, TRPV2, TRPV4, TRPM4, TRPM6, and TRPM7 are abundantly expressed, as assessed by RT-PCR. Moreover, all the TRP channel transcripts have been detected by qPCR in isolated mouse cardiac fibroblasts [37]. Although the expression of TRP channels can be readily detected by RT-PCR, only a few TRP channels can be functionally detected by patch-clamp current recording, as summarized in Table 1.
The first TRP channel current recorded in cardiac fibroblasts was the TRPM7 current recorded in rat cardiac fibroblasts [72]. TRPM7 currents can also be readily recorded in other fibroblasts, such as in human and mouse cardiac fibroblasts [37]. The TRPM7-like currents recorded in mouse and human fibroblasts exhibit similar biophysical and pharmacological properties to those of the heterologously-expressed TRPM7 currents [137,149,161,162], including sensitivity to external acidic pHs, and a distinct response to 2-APB in comparison to TRPM6. TRPC3, TRPC6, or TRPC3/6 heteromer-like currents have been reported in rat ventricular fibroblasts [69], and TRPC3-like currents have also been recorded in rat atrial fibroblasts [70]. Similarly, a TRPM2-like current has been recorded in rat cardiac fibroblasts after 24 hours of H2O2 treatment, suggesting hypoxia up-regulates TRPM2 in rat cardiac fibroblasts [59]. The TRPM2-like current can be inhibited by Clotrimazole, and reduced by TRPM2 siRNA treatment. Using 4α-12,13-didecanoate (4α-PDD) as an activator, TRPV4-like currents have been recorded in rat cardiac fibroblasts. The 4α-PDD-induced TRPV4 current is blocked by ruthenium red, and is reduced by TRPV4 siRNA [71]. Thus, many different TRP channels are expressed in cardiac fibroblasts, and recent work is adding to a growing body of evidence that TRP channels in cardiac cells, including myocytes and fibroblasts, play an essential role in heart diseases [159,163]. In this review, we focus on summarizing the recent advances regarding the role of TRP channels in mediating Ca2+ signaling in cardiac fibroblasts, and their potential roles in cardiac fibrosis and fibrosis-associated heart diseases [37,39,70,164,165] (Table 2).

3.4. Ca2+ Signaling Mediated by TRP Channels in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases

3.4.1. TRPC Channels

The role of TRPC channels in hypertrophy and heart failure has been extensively studied by the systematic or myocardial knockdown, overexpression, and knockout of TRPC channels, as previously reviewed [159,163,165,201]. Most of the TRPC channels, TRPC1, TRPC3, TRPC6, TRPC7, and the heterotetrameric channel complexes TRPC3/6/7, TRPC1/4/5, and TRPC1/TRPC4 [202], have been demonstrated to be important mediators of pathological hypertrophy, and may serve as therapeutic targets [201,203]. Indeed, the TRPC3 blocker Pyrazole-3 (Pyr3) [166], the combined TRPC3 and TRPC6 channel blockers GSK2332255B and GSK2833503A [204], and the TRPC6 channel blocker BI-749327 [38], are effective at attenuating pathological remodeling and improving heart function. Although these studies provide strong evidence that TRPC channels in cardiac myocytes play an essential role in mediating hypertrophic remodeling, it remains unclear whether TRPC channels in cardiac fibroblasts contribute to pathological remodeling.
Despite knowledge of the role of TRPC channels in fibroblasts in hypertrophic remodeling being limited, it has been shown that the pathological stimuli AngII and ET1 can indeed induce Ca2+ entry through TRPC3 [70] and TRPC6 [160], respectively. Moreover, nucleotides released during ischemic injury activate P2Y2 and P2Y4 receptors to cause Ca2+ release and subsequent Ca2+ entry via TRPC channel activation, which leads to fibroblast differentiation [205]. Nonetheless, the role of TRPC3 and TRPC6 in fibrogenesis has been investigated in great detail recently [70,166].
TRPC3 and Fibrosis-Associated Arrhythmia and Heart Failure
TRPC3 channels have been shown to play a role in cardiac fibrosis and fibrosis-associated heart diseases, such as atrial fibrillation (AF) [70] and heart failure induced by pressure overload [166,167]. It has been recently demonstrated that TRPC3, independent of TRPC6, mediates pressure-overload-induced maladaptive cardiac fibrosis [168,169]. Pharmacological inhibition of TRPC3 suppresses the fibrotic response in human cardiac myocytes and fibroblasts [169]. Administration of the combined TRPC3 and TRPC6 blocker GSK503A to pressure-overload mice or rats also results in antifibrotic effects [204]. In pressure-overload mouse hearts, the inhibition of TRPC3 reduces RhoA-mediated maladaptive fibrosis [169]. It was proposed that TRPC3-mediated intracellular Ca2+ signaling activates PKC, which then phosphorylates p47phox and activates NADPH oxidase 2 (Nox2) to generate ROS, thereby initiating the RhoA signaling pathway and ECM production [167,168,169].
TRPC3-mediated Ca2+ signaling in fibroblasts is also involved in arrhythmogenesis [70]. TRPC3 is upregulated in the atria of atrial fibrillation (AF) patients, and in atrial fibrillation goat, and dog models [70]. The mechanism by which atrial fibrillation increases TRPC3 expression is mediated by NFAT-induced downregulation of microRNA-26 [70]. Enhanced TRPC3 expression increases fibroblast proliferation and differentiation, likely by controlling Ca2+ influx, which activates extracellular signal-regulated kinase (ERK1/2) signaling [70]. TRPC3-mediated fibroblast proliferation and differentiation through the ERK signaling pathway has also been reported in other studies [170]. In vivo administration of the TRPC3 blocker Pyr3 reduces ECM protein expression and suppresses development of the AF substrate in the electrically-maintained dog model of atrial fibrillation [70]. Thus, reducing TRPC3-mediated Ca2+ signaling appears to reduce the susceptibility to AF, perhaps by reducing fibroblast cell proliferation. Overall, the data suggest that TRPC3 is likely a potential therapeutic target for fibrosis-associated atrial fibrillation [70].
TRPC6-Mediated Fibroblast Differentiation and Heart Failure
TRPC6 has been reported to regulate fibroblast differentiation induced by ET1 in rat ventricular fibroblasts [160]. TRPC6-mediated Ca2+ influx by ET1 in neonatal rat ventricular fibroblasts activates NFAT, which acts as a negative regulator of ET1-induced fibroblast differentiation [160]. In contrast, other studies report that TRPC6 instead promotes fibroblast proliferation and differentiation [171]. It has also been demonstrated that TRPC6-mediated Ca2+ signaling is required for human cardiac fibroblast proliferation induced by AngII and OAG (1-oleoyl-2-acetyl-sn-glycerol); the latter is a diacylglycerol (DAG) analogue [171]. Consistent with a positive role for TRPC6 in fibroblast differentiation, Davis and colleagues have demonstrated that TRPC6 is necessary and sufficient for fibroblast differentiation induced by AngII [164]. It has been shown that TRPC6 is upregulated by TGF-β1 and AngII via the p38 MAPK (mitogen-activated protein kinase) serum response factor [164]. Activation of TRPC6 stimulates the calcineurin/NFAT pathway to induce fibroblast differentiation [164]. Fibroblasts lacking TRPC6 (TRPC6-/-) were not able to differentiate into myofibroblasts in response to TGF-β1 stimulation [164]. Also, mice without TRPC6 displayed impaired dermal and heart-wound healing function [164]. Consistent with the role of TRPC6 in promoting fibroblast differentiation, silencing TRPC6 with a siRNA approach attenuated the TGF-β1-mediated upregulation of α-SMA in the human right ventricle [172]. Moreover, administration of the TRPC6 antagonist (BI-749327) to mice subjected to pressure-overload inhibited the profibrotic gene expression, reduced cardiac fibrosis, and improved heart function [38], suggesting that inhibition of TRPC6 in fibroblasts at least partially mediates the protective effects in the attenuation of pressure-overload-induced cardiac remodeling.

3.4.2. TRPV Channels in Fibroblasts and Cardiac Fibrosis

Several TRPV channels have been implicated in cardiac fibrosis, including TRPV1, TRPV2, TRPV3, and TRPV4. We summarize their putative roles below.
TRPV1 and Cardiac Fibrosis
TRPV1 is predominately expressed in peripheral sensory neurons and is widespread in the cardiovascular system [206]. The effects of TRPV1 in cardiac fibrosis have been controversial. Some studies demonstrate that deletion of TRPV1 results in stimulation of the TGF-β1 and SMAD2 signaling pathway, and therefore, significantly increases fibrosis in a myocardial injury model [207]. In contrast, another study reported that the activation of TRPV1 by capsaicin blunts pressure-overload-induced hypertrophy and fibrosis [173]. It has also been shown that capsaicin reduces fibroblast proliferation induced by AngII in vitro [173]. Moreover, overexpression of TRPV1 in transgenic mice attenuates isoproterenol-induced myocardial fibrosis [174], and the activation of TRPV1 has also been shown to be protective in a myocardial injury model [175,176,177]. On the contrary, other studies demonstrate that administration of a TRPV1 antagonist BCTC (4-(3-Chloro-2-pyridinyl)-N-[4-(1,1-dimethylethyl)phenyl]-1-piperazinecarboxamide) prevents the loss of heart function, and protects the heart from fibrosis in a pressure-overload mouse model [208]. Mice lacking TRPV1 are also protected from pressure-overload-induced cardiac hypertrophy [209]. The cause of these discrepancies in different studies regarding the role of TRPV1 in fibrogenesis is unclear, but it could be attributed to the expression of TRPV1 in different cell types and the channel’s involvement in different signaling pathways.
TRPV2 and Cardiac Fibrosis
TRPV2 has been shown to play a role in Ca2+-induced myocyte degeneration in dilated cardiomyopathy [178]. A knockout of TRPV2 protects hearts against pressure-overload-induced hypertrophy, but produces no protection against AngII or β-adrenergic activation-induced hypertrophy, indicating that TRPV2 regulates cardiac hypertrophy through stretch activation [179]. A knockout of TRPV2 in mice also reduces age-related fibrosis and hypertrophy [210]. Application of the TRPV2 blocker Tranilast blunts the hypertrophic and fibrotic response to pressure-overload-induced hypertrophy within four weeks [180]. Although TRPV2 has been shown to be located intracellularly and translocates to the sarcolemma under stress, the effect of Tranilast does not seem to be mediated through its effects on myocytes [180]. TRPV2 is also expressed in other cell types, such as macrophages, and in fact, TRPV2 knockout improves heart performance after myocardial infarction due to attenuated activity of peri-infarct macrophages [181]. TRPV2 is also highly expressed in cardiac fibroblasts and has been shown to regulate dermal fibroblast differentiation [211]. Thus, it is plausible that TRPV2-mediated Ca2+ signaling in cardiac fibroblasts regulates fibroblast differentiation and cardiac fibrosis. Further studies are required to dissect the different roles of TRPV2 during heart injury.
TRPV3 and Cardiac Fibrosis
TRPV3 is expressed in the peripheral neurons and various other cell types. It has recently been found that TRPV3 is also expressed in cardiac myocytes [182] and fibroblasts [183], as detected by PCR and western blotting (WB). In rat cardiac fibroblasts, the activation of TRPV3 by Carvacrol increases cell proliferation and upregulates the expression of collagen and TGF-β1. In a pressure-overload hypertrophy rat model, Carvacrol activation of TRPV3 exacerbates heart function and increases fibrosis [182,183]. These results suggest that TRPV3 could be involved in TGF-β1-induced fibroblast proliferation during cardiac fibrosis, but additional experiments with knockout mice are required to conclusively define the channel’s role.
TRPV4 and Cardiac Fibrosis
TRPV4 has been shown to regulate fibroblasts’ differentiation into myofibroblasts by integrating TGF-β1 signals and mechanical stimulation [39,184]. In cultured rat fibroblasts, TRPV4 agonists elicit a substantial Ca2+ influx. The application of a TRPV4 antagonist and an shRNA knockdown of TRPV4 inhibits TGF-β1-induced myofibroblast differentiation, whereas TGF-β1-treated fibroblasts exhibit enhanced TRPV4 expression and increased Ca2+ influx [39]. In a lung fibrosis mouse model, it has been shown that deletion of TRPV4 protects mice from lung fibrosis [212]. Similarly, in a myocardial infarction model, TRPV4 deletion in mice also reduces fibrosis and protects the heart from pathological remodeling [185]. These studies highlight an important role for TRPV4 in cardiac fibrosis, making it a potentially attractive therapeutic target.

3.4.3. TRPM Channels and Cardiac Fibrosis

Several TRPM channels have been shown to play important physiological or pathological roles in the heart. TRPM2 has been found to be involved in ischemic cardiomyopathy, with protective roles reported by some studies [186,187], and exacerbated roles reported by others [188,189]. In cardiac fibroblasts, TRPM2 current can be induced in rat fibroblasts after treatment with H2O2 [59]. However, whether TRPM2 channel activity influences fibroblast function remains unknown. The role of TRPM4 in cardiac fibroblasts has not been evaluated, although TRPM4 channel function has been shown to be essential for cardiac conduction, and the dysfunction of TRPM4 is associated with conduction defects and arrhythmia [213,214,215,216].
Physiological Function of TRPM7 in the Heart
TRPM7 is essential for embryogenesis [217], organogenesis [218], and cardiogenesis [190,191]. The global deletion of TRPM7 in mice results in embryonic lethality before embryonic day 7 (E7). Myocyte-specific deletion of TRPM7 before E9 results in impaired compact myocardium development with consequential congestive heart failure and embryonic death by E11.5 [190]. However, myocyte deletion of TRPM7 at about E13 produces viable mice with normal adult ventricular sizes and heart function [190], indicating that TRPM7 is dispensable for the adult mice [190]. However, if TRPM7 is deleted in myocytes between day E9.5 and E12.5, 50% of mice have normal heart function, and another 50% of mice exhibit penetrant adult cardiomyopathy characterized by ventricular dysfunction, hypertrophy, fibrosis, disputed atrioventricular conduction, dispersed ventricular repolarization, and ventricular arrhythmia [190]. These results indicate that TRPM7 deletion in myocytes during the developmental stage (E9.5 to E12.5) can cause impaired adult heart function, whereas deletion of TRPM7 after embryonic day 13 does not produce any effects on normal heart function [190]. Interestingly, a recent study by Rios and colleagues reported that TRPM7-deficient mice with kinase domain deletion (TRPM7+/Δkinase) exhibit cardiac hypertrophy, fibrosis, and inflammation [219]. It was suggested that TRPM7 plays both anti-fibrotic and anti-inflammatory roles [219]. However, since TRPM7 channel activity was impaired in the TRPM7+/Δkinase mice [220], it is plausible that the disrupted TRPM7 channel function during the developmental stage (E9.5 to E12.5) may have contributed to the impaired heart function, resulting in the observed cardiac hypertrophy, inflammation, and fibrosis in the TRPM7+/Δkinase mice [219].
Expression of TRPM7 in Cardiac Fibroblasts
TRPM7 is also highly expressed in cardiac fibroblasts. Endogenous TRPM7-like currents were initially recorded in rat cardiac fibroblasts [72], and were readily recorded in mouse and human fibroblasts [37,221,222]. TRPM7-like current is significantly upregulated in fibroblasts from atrial fibrillation patients [37]. Although TRPM6 expression was found to be upregulated in the right atria of atrial fibrillation patients in comparison with sinus rhythm patients [223], the endogenous currents recorded in human atrial fibroblasts are encoded by TRPM7, as TRPM7 siRNA eliminates the majority of TRPM7-like currents. TRPM6 and TRPM7 can form heteromeric channels [137,149,161]. The homomeric TRPM6 and TRPM7 channels and heteromeric TRPM6/7 channels share similar biophysical properties, such as current-voltage (I-V) relation, the potentiation for inward current by externally low pH, and regulation by PIP2, but they have distinct single-channel conductances and pharmacological sensitivities to low and high concentrations of 2-APB [137,149,161,224]. The endogenous TRPM7-like currents recorded in human atrial fibroblasts display a similar single-channel conductance and 2-APB sensitivity to that of TRPM7 currents recorded in the over-expression system, indicating that the functional current in human atrial fibroblasts is encoded by TRPM7. It will be of interest to investigate whether TRPM6 contributes to functional expression of TRPM7-like currents in cardiac fibroblasts.
TRPM7 and Fibroblast Differentiation
TRPM7-mediated Ca2+ entry is significantly enhanced in fibroblasts from AF patients. Knockdown of TRPM7 inhibits TGF-β1-induced fibroblast proliferation, differentiation, and collagen production [37], whereas upregulation of TRPM7 by TGF-β1 in cultured human atrial fibroblasts or mouse cardiac fibroblasts enhances fibroblast proliferation, differentiation, and ECM production [37]. It appears that TRPM7-mediated Ca2+ signaling is essential for TGF-β1-induced fibroblast proliferation, differentiation, and the fibrogenesis cascade [37]. A role of TRPM7 in mediating the fibrogenesis cascade has also been observed in rat cardiac fibroblasts [193,194,195,196]. Moreover, TRPM7 is also suggested to be involved in fibrogenesis in rat sinus nodes [192], and in fibrosis induced by isoproterenol and oxidative stress [195]. Thus, it seems that TRPM7 may serve as a potential target for fibrosis-associated heart diseases.

3.4.4. TRPA1 in Cardiac Fibroblasts

TRPA1 is expressed in cardiac myocytes [197,198] and fibroblasts [199]. Activation of TRPA1 elicits Ca2+ influx in primary cultured human ventricular cardiac fibroblasts [73]. TRPA1 selective inhibitors HC-030031 (HC) and TCS-5861528 (TCS) ameliorate pressure-overload-induced cardiac hypertrophy by negatively regulating Ca2+/calmodulin-dependent protein kinase II (CaMKII) and calcineurin signaling pathways [200]. The knockout TRPA1 increases survival rate, reduces fibrosis, and enhances heart performance after myocardial infarction [199]. Moreover, in primary cultured cardiac fibroblasts, an overexpression of TRPA1 potentiates, whereas a knockdown TRPA1 attenuates, TGF-β1-induced fibroblast differentiation. Finally, it appears that TRPA1 regulates fibroblast differentiation through the calcineurin-dependent NFAT3 activation pathway [200].

3.5. Ca2+ Efflux and Fibroblast Function

In contrast to the sarcoplasmic reticulum (SR) Ca2+-ATPase (SERCA) and the plasma membrane Na+/Ca2+ exchanger (NCX), the two major components that extrude cytosolic Ca2+ to the SR/ER and outside of the cell, the plasma membrane Ca2+-ATPase (PMCA1,2,3,4) was considered to be relatively insignificant in maintaining Ca2+ homeostasis in the heart [225,226]. In recent years, however, it has been demonstrated that PMCA4 plays a significant role in the regulation of signal transduction in the heart.
Among the four PMCAs, the expression of PMCAs 1, 3, and 4 has been detected at the mRNA level in human cardiac fibroblasts [46]. Recently, it has been shown that the deletion of PMCA4 in mouse fibroblasts, but not in myocytes, protects the heart against hypertensive hypertrophy [227]. The basal Ca2+ level in PMCA4-KO fibroblasts is 25% higher than that in control fibroblasts. PMCA4 ablation also increases secreted, frizzled-related protein 2 (sFRP2) expression, which inhibits the hypertrophic response in myocytes [227]. Moreover, the PMCA4 inhibitor aurintricarboxylic acid (ATA) has been reported to inhibit and reverse cardiac hypertrophy induced by pressure-overload in mice [227]. These results indicate that regulating Ca2+ homeostasis in fibroblasts can also mitigate fibrosis-associated heart diseases.
NCX and SERCA are essential to maintaining Ca2+ homeostasis in many cell types, such as cardiac myocytes. In cardiac fibroblasts, the mRNA expression of NCX and SERCA has also been detected in human ventricular fibroblasts [46]. It was suggested that NCX plays a role in Ca2+ homeostasis in fibroblasts [46], based on the effects of the NCX inhibitor, Ni2+, and extracellular Na+ replacement, on Ca2+ oscillations in human ventricular fibroblasts. However, the detailed characterization of NCX and SERCA in cardiac fibroblasts, and their roles in fibroblast function and in cardiac fibrosis, still remain largely unknown.

4. Conclusions and Future Perspectives

Fibrosis is a hallmark feature of most heart diseases. There are a multitude of signaling pathways, bioactive molecules, and various cell types that are involved in the cardiac fibrogenesis cascade. As the major cell type in the fibrogenesis cascade, fibroblasts are the receivers of various pathological stimuli and the producers of extracellular matrix proteins. However, the Ca2+ signaling mechanisms controlling cardiac fibroblast functions are not fully understood. Nevertheless, the advances in recent years have helped to shape our understanding of the complex nature of Ca2+ signaling mechanisms in fibroblasts, while TRP channels with their unique features have emerged as the most important ion channels that mediate Ca2+ signals in cardiac fibroblasts. Evidence suggests that many TRP channels have the potential to be therapeutic targets for drug intervention. However, more work is necessary to understand TRP channels’ functions in the heart. As TRP channels are widely expressed in different cell types, future investigations using cell type-specific deletions of TRP channel genes will provide more precise information regarding their specific roles in cardiac fibroblasts. Moreover, as in vitro fibroblast experiments are an important part of cardiac fibroblast research, finding the experimental conditions that closely mimic the in vivo properties of fibroblasts and myofibroblasts will be critical. Finally, since fibrogenesis involves multiple cell types in the heart, gaining a better understanding of the fibroblast differentiation process, and how Ca2+ signaling influences the interactions of fibroblasts and myofibroblasts with myocytes and other cell types, is an important challenge whose outcomes will provide new insights into the therapeutic potential of TRP channels in fibrosis-associated heart diseases.

Author Contributions

L.Y. conceived the concept of the review. L.W.R. and L.Y. wrote the manuscript. J.F. and M.K.A. designed and formatted the figures. J.F., M.K.A., A.S.Y., B.T.L., L.W.R., and L.Y. read, edited, and reviewed the manuscript.

Acknowledgments

This work was generously supported by the National Institutes of Health, National Heart, Lung and Blood Institute (NHLBI)—R01 HL147350 to L.W.R. and L.Y., and NIH 2R01HL078960 to L.Y.; the American Heart Association (AHA)—19TPA34890022 to L.Y.; and the AHA predoctoral fellowship (AHA 17PRE33410019) to A.S.Y.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Weber, K. Cardiac interstitium. In Heart Failure; Poole-Wilson, P., Colucci, W., Massie, B., Chatterjee, K., Coats, A., Eds.; Churchill Livingstone: New York, NY, USA, 1997; pp. 13–31. [Google Scholar]
  2. Gourdie, R.G.; Dimmeler, S.; Kohl, P. Novel therapeutic strategies targeting fibroblasts and fibrosis in heart disease. Nat. Rev. Drug Discov. 2016, 15, 620–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Travers, J.G.; Kamal, F.A.; Robbins, J.; Yutzey, K.E.; Blaxall, B.C. Cardiac Fibrosis: The Fibroblast Awakens. Circ. Res. 2016, 118, 1021–1040. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, Z.G.; Yuan, Y.P.; Wu, H.M.; Zhang, X.; Tang, Q.Z. Cardiac fibrosis: New insights into the pathogenesis. Int. J. Biol. Sci. 2018, 14, 1645–1657. [Google Scholar] [CrossRef] [PubMed]
  5. Nguyen, M.N.; Kiriazis, H.; Gao, X.M.; Du, X.J. Cardiac Fibrosis and Arrhythmogenesis. Compr. Physiol. 2017, 7, 1009–1049. [Google Scholar] [CrossRef] [PubMed]
  6. Weber, K.T.; Sun, Y.; Diez, J. Fibrosis: A Living Tissue and the Infarcted Heart. J. Am. Coll. Cardiol. 2008, 52, 2029–2031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. de Bakker, J.M.; van Capelle, F.J.; Janse, M.J.; Tasseron, S.; Vermeulen, J.T.; de Jonge, N.; Lahpor, J.R. Fractionated electrograms in dilated cardiomyopathy: Origin and relation to abnormal conduction. J. Am. Coll. Cardiol. 1996, 27, 1071–1078. [Google Scholar] [CrossRef] [Green Version]
  8. Burstein, B.; Nattel, S. Atrial Fibrosis: Mechanisms and Clinical Relevance in Atrial Fibrillation. J. Am. Coll. Cardiol. 2008, 51, 802–809. [Google Scholar] [CrossRef] [Green Version]
  9. Yue, L.; Xie, J.; Nattel, S. Molecular determinants of cardiac fibroblast electrical function and therapeutic implications for atrial fibrillation. Cardiovasc Res. 2011, 89, 744–753. [Google Scholar] [CrossRef]
  10. Khan, R.; Sheppard, R. Fibrosis in heart disease: Understanding the role of transforming growth factor-beta in cardiomyopathy, valvular disease and arrhythmia. Immunology 2006, 118, 10–24. [Google Scholar] [CrossRef]
  11. Weber, K.T. Are myocardial fibrosis and diastolic dysfunction reversible in hypertensive heart disease? Congest. Heart Fail. 2005, 11, 322–324. [Google Scholar] [CrossRef]
  12. Kania, G.; Blyszczuk, P.; Eriksson, U. Mechanisms of cardiac fibrosis in inflammatory heart disease. Trends Cardiovasc. Med. 2009, 19, 247–252. [Google Scholar] [CrossRef] [PubMed]
  13. Biernacka, A.; Frangogiannis, N.G. Aging and Cardiac Fibrosis. Aging Dis. 2011, 2, 158–173. [Google Scholar] [PubMed]
  14. Pellman, J.; Zhang, J.; Sheikh, F. Myocyte-fibroblast communication in cardiac fibrosis and arrhythmias: Mechanisms and model systems. J. Mol. Cell. Cardiol. 2016, 94, 22–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Gonzalez, A.; Schelbert, E.B.; Diez, J.; Butler, J. Myocardial Interstitial Fibrosis in Heart Failure: Biological and Translational Perspectives. J. Am. Coll. Cardiol. 2018, 71, 1696–1706. [Google Scholar] [CrossRef] [PubMed]
  16. Segura, A.M.; Frazier, O.H.; Buja, L.M. Fibrosis and heart failure. Heart Fail. Rev. 2014, 19, 173–185. [Google Scholar] [CrossRef] [PubMed]
  17. Sun, Y.; Kiani, M.F.; Postlethwaite, A.E.; Weber, K.T. Infarct scar as living tissue. Basic Res. Cardiol. 2002, 97, 343–347. [Google Scholar] [CrossRef]
  18. Marijianowski, M.M.; Teeling, P.; Mann, J.; Becker, A.E. Dilated cardiomyopathy is associated with an increase in the type I/type III collagen ratio: A quantitative assessment. J. Am. Coll. Cardiol. 1995, 25, 1263–1272. [Google Scholar] [CrossRef] [Green Version]
  19. Sanderson, J.E.; Lai, K.B.; Shum, I.O.; Wei, S.; Chow, L.T. Transforming growth factor-beta(1) expression in dilated cardiomyopathy. Heart 2001, 86, 701–708. [Google Scholar] [CrossRef]
  20. Kuhl, U.; Noutsias, M.; Schultheiss, H.P. Immunohistochemistry in dilated cardiomyopathy. Eur. Heart J. 1995, 16 (Suppl. O), 100–106. [Google Scholar] [CrossRef]
  21. Dixon, I.M.; Hao, J.; Reid, N.L.; Roth, J.C. Effect of chronic AT(1) receptor blockade on cardiac Smad overexpression in hereditary cardiomyopathic hamsters. Cardiovasc. Res. 2000, 46, 286–297. [Google Scholar] [CrossRef]
  22. Pearlman, E.S.; Weber, K.T.; Janicki, J.S.; Pietra, G.G.; Fishman, A.P. Muscle fiber orientation and connective tissue content in the hypertrophied human heart. Lab. Investig. J. Tech. Methods Pathol. 1982, 46, 158–164. [Google Scholar]
  23. Brilla, C.G. Aldosterone and myocardial fibrosis in heart failure. Herz 2000, 25, 299–306. [Google Scholar] [CrossRef] [PubMed]
  24. Bacmeister, L.; Schwarzl, M.; Warnke, S.; Stoffers, B.; Blankenberg, S.; Westermann, D.; Lindner, D. Inflammation and fibrosis in murine models of heart failure. Basic Res. Cardiol. 2019, 114, 19. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, X.; Chen, B.; Liu, T.; Chen, X. Reversal of myofibroblast differentiation: A review. Eur. J. Pharmacol. 2014, 734, 83–90. [Google Scholar] [CrossRef] [PubMed]
  26. Petrov, V.V.; Fagard, R.H.; Lijnen, P.J. Stimulation of Collagen Production by Transforming Growth Factor-β1 During Differentiation of Cardiac Fibroblasts to Myofibroblasts. Hypertension 2002, 39, 258–263. [Google Scholar] [CrossRef] [PubMed]
  27. Porter, K.E.; Turner, N.A. Cardiac fibroblasts: At the heart of myocardial remodeling. Pharmacol. Ther. 2009, 123, 255–278. [Google Scholar] [CrossRef] [PubMed]
  28. Long, C.S.; Brown, R.D. The cardiac fibroblast, another therapeutic target for mending the broken heart? J. Mol. Cell. Cardiol. 2002, 34, 1273–1278. [Google Scholar] [CrossRef] [PubMed]
  29. Weber, K.T.; Brilla, C.G. Pathological hypertrophy and cardiac interstitium. Fibrosis and renin- angiotensin-aldosterone system. Circulation 1991, 83, 1849–1865. [Google Scholar] [CrossRef]
  30. Brilla, C.G.; Janicki, J.S.; Weber, K.T. Cardioreparative effects of lisinopril in rats with genetic hypertension and left ventricular hypertrophy. Circulation 1991, 83, 1771–1779. [Google Scholar] [CrossRef]
  31. Brilla, C.G.; Pick, R.; Tan, L.B.; Janicki, J.S.; Weber, K.T. Remodeling of the rat right and left ventricles in experimental hypertension. Circ. Res. 1990, 67, 1355–1364. [Google Scholar] [CrossRef]
  32. Smits, J.F.; Cleutjens, J.P.; van Krimpen, C.; Schoemaker, R.G.; Daemen, M.J. Cardiac remodeling in hypertension and following myocardial infarction: Effects of arteriolar vasodilators. Basic Res. Cardiol. 1991, 86, 133–139. [Google Scholar] [PubMed]
  33. Weber, K.T.; Sun, Y.; Tyagi, S.C.; Cleutjens, J.P. Collagen network of the myocardium: Function, structural remodeling and regulatory mechanisms. J. Mol. Cell. Cardiol. 1994, 26, 279–292. [Google Scholar] [CrossRef] [PubMed]
  34. Swynghedauw, B. Molecular mechanisms of myocardial remodeling. Physiol. Rev. 1999, 79, 215–262. [Google Scholar] [CrossRef] [PubMed]
  35. Leask, A. Getting to the heart of the matter: New insights into cardiac fibrosis. Circulation Res. 2015, 116, 1269–1276. [Google Scholar] [CrossRef]
  36. Yue, Z.; Zhang, Y.; Xie, J.; Jiang, J.; Yue, L. Transient receptor potential (TRP) channels and cardiac fibrosis. Curr. Top. Med. Chem. 2013, 13, 270–282. [Google Scholar] [CrossRef] [PubMed]
  37. Du, J.; Xie, J.; Zhang, Z.; Tsujikawa, H.; Fusco, D.; Silverman, D.; Liang, B.; Yue, L. TRPM7-mediated Ca2+ signals confer fibrogenesis in human atrial fibrillation. Circ. Res. 2010, 106, 992–1003. [Google Scholar] [CrossRef] [PubMed]
  38. Lin, B.L.; Matera, D.; Doerner, J.F.; Zheng, N.; Del Camino, D.; Mishra, S.; Bian, H.; Zeveleva, S.; Zhen, X.; Blair, N.T.; et al. In vivo selective inhibition of TRPC6 by antagonist BI 749327 ameliorates fibrosis and dysfunction in cardiac and renal disease. Proc. Natl. Acad. Sci. USA 2019, 116, 10156–10161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Adapala, R.K.; Thoppil, R.J.; Luther, D.J.; Paruchuri, S.; Meszaros, J.G.; Chilian, W.M.; Thodeti, C.K. TRPV4 channels mediate cardiac fibroblast differentiation by integrating mechanical and soluble signals. J. Mol. Cell. Cardiol. 2013, 54, 45–52. [Google Scholar] [CrossRef] [PubMed]
  40. Kiseleva, I.S.; Kamkin, A.G.; Kirkheis, R.; Kositskii, G.I. Intercellular electrotonic interactions in the cardiac sinus node in the frog. Doklady Akademii Nauk SSSR 1987, 292, 1502–1505. [Google Scholar] [PubMed]
  41. Kamkin, A.; Kiseleva, I.; Wagner, K.D.; Pylaev, A.; Leiterer, K.P.; Theres, H.; Scholz, H.; Gunther, J.; Isenberg, G. A possible role for atrial fibroblasts in postinfarction bradycardia. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, H842–H849. [Google Scholar] [CrossRef] [Green Version]
  42. Kamkin, A.; Kiseleva, I.; Wagner, K.D.; Lammerich, A.; Bohm, J.; Persson, P.B.; Gunther, J. Mechanically induced potentials in fibroblasts from human right atrium. Exp. Physiol. 1999, 84, 347–356. [Google Scholar] [CrossRef]
  43. Kamkin, A.; Kiseleva, I.; Isenberg, G. Activation and inactivation of a non-selective cation conductance by local mechanical deformation of acutely isolated cardiac fibroblasts. Cardiovasc. Res. 2003, 57, 793–803. [Google Scholar] [CrossRef]
  44. Aguilar, M.; Qi, X.Y.; Huang, H.; Comtois, P.; Nattel, S. Fibroblast electrical remodeling in heart failure and potential effects on atrial fibrillation. Biophys. J. 2014, 107, 2444–2455. [Google Scholar] [CrossRef] [PubMed]
  45. Kohl, P.; Noble, D. Mechanosensitive connective tissue: Potential influence on heart rhythm. Cardiovasc. Res. 1996, 32, 62–68. [Google Scholar] [CrossRef]
  46. Chen, J.B.; Tao, R.; Sun, H.Y.; Tse, H.F.; Lau, C.P.; Li, G.R. Multiple Ca2+ signaling pathways regulate intracellular Ca2+ activity in human cardiac fibroblasts. J. Cell. Physiol. 2010, 223, 68–75. [Google Scholar] [CrossRef] [PubMed]
  47. Walsh, K.B.; Zhang, J. Neonatal rat cardiac fibroblasts express three types of voltage-gated K+ channels: Regulation of a transient outward current by protein kinase C. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H1010–H1017. [Google Scholar] [CrossRef]
  48. Shibukawa, Y.; Chilton, E.L.; Maccannell, K.A.; Clark, R.B.; Giles, W.R. K+ currents activated by depolarization in cardiac fibroblasts. Biophys. J. 2005, 88, 3924–3935. [Google Scholar] [CrossRef]
  49. Dawson, K.; Wu, C.T.; Qi, X.Y.; Nattel, S. Congestive heart failure effects on atrial fibroblast phenotype: Differences between freshly-isolated and cultured cells. PLoS ONE 2012, 7, e52032. [Google Scholar] [CrossRef]
  50. Wu, C.T.; Qi, X.Y.; Huang, H.; Naud, P.; Dawson, K.; Yeh, Y.H.; Harada, M.; Kuo, C.T.; Nattel, S. Disease and region-related cardiac fibroblast potassium current variations and potential functional significance. Cardiovasc. Res. 2014, 102, 487–496. [Google Scholar] [CrossRef] [Green Version]
  51. Li, G.R.; Sun, H.Y.; Chen, J.B.; Zhou, Y.; Tse, H.F.; Lau, C.P. Characterization of multiple ion channels in cultured human cardiac fibroblasts. PLoS ONE 2009, 4, e7307. [Google Scholar] [CrossRef]
  52. Chilton, L.; Ohya, S.; Freed, D.; George, E.; Drobic, V.; Shibukawa, Y.; MacCannell, K.A.; Imaizumi, Y.; Clark, R.B.; Dixon, I.M.C.; et al. K+ currents regulate the resting membrane potential, proliferation, and contractile responses in ventricular fibroblasts and myofibroblasts. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H2931–H2939. [Google Scholar] [CrossRef]
  53. Qi, X.Y.; Huang, H.; Ordog, B.; Luo, X.; Naud, P.; Sun, Y.; Wu, C.T.; Dawson, K.; Tadevosyan, A.; Chen, Y.; et al. Fibroblast inward-rectifier potassium current upregulation in profibrillatory atrial remodeling. Circ. Res. 2015, 116, 836–845. [Google Scholar] [CrossRef]
  54. Benamer, N.; Moha Ou Maati, H.; Demolombe, S.; Cantereau, A.; Delwail, A.; Bois, P.; Bescond, J.; Faivre, J.F. Molecular and functional characterization of a new potassium conductance in mouse ventricular fibroblasts. J. Mol. Cell. Cardiol. 2009, 46, 508–517. [Google Scholar] [CrossRef]
  55. Pertiwi, K.R.; Hillman, R.M.; Scott, C.A.; Chilton, E.L. Ischemia Reperfusion Injury Produces, and Ischemic Preconditioning Prevents, Rat Cardiac Fibroblast Differentiation: Role of KATP Channels. J. Cardiovasc. Dev. Dis. 2019, 6. [Google Scholar] [CrossRef]
  56. Benamer, N.; Vasquez, C.; Mahoney, V.M.; Steinhardt, M.J.; Coetzee, W.A.; Morley, G.E. Fibroblast KATP currents modulate myocyte electrophysiology in infarcted hearts. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H1231–H1239. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, Y.J.; Sung, R.J.; Lin, M.W.; Wu, S.N. Contribution of BK(Ca)-channel activity in human cardiac fibroblasts to electrical coupling of cardiomyocytes-fibroblasts. J. Membr. Biol. 2006, 213, 175–185. [Google Scholar] [CrossRef]
  58. He, M.L.; Liu, W.J.; Sun, H.Y.; Wu, W.; Liu, J.; Tse, H.F.; Lau, C.P.; Li, G.R. Effects of ion channels on proliferation in cultured human cardiac fibroblasts. J. Mol. Cell. Cardiol. 2011, 51, 198–206. [Google Scholar] [CrossRef]
  59. Takahashi, K.; Sakamoto, K.; Kimura, J. Hypoxic stress induces transient receptor potential melastatin 2 (TRPM2) channel expression in adult rat cardiac fibroblasts. J. Pharmacol. Sci. 2012, 118, 186–197. [Google Scholar] [CrossRef]
  60. Stockbridge, L.L.; French, A.S. Stretch-activated cation channels in human fibroblasts. Biophys. J. 1988, 54, 187–190. [Google Scholar] [CrossRef] [Green Version]
  61. Chatelier, A.; Mercier, A.; Tremblier, B.; Theriault, O.; Moubarak, M.; Benamer, N.; Corbi, P.; Bois, P.; Chahine, M.; Faivre, J.F. A distinct de novo expression of Nav1.5 sodium channels in human atrial fibroblasts differentiated into myofibroblasts. J. Physiol. 2012, 590, 4307–4319. [Google Scholar] [CrossRef]
  62. Koivumaki, J.T.; Clark, R.B.; Belke, D.; Kondo, C.; Fedak, P.W.; Maleckar, M.M.; Giles, W.R. Na(+) current expression in human atrial myofibroblasts: Identity and functional roles. Front. Physiol. 2014, 5, 275. [Google Scholar] [CrossRef]
  63. Poulet, C.; Kunzel, S.; Buttner, E.; Lindner, D.; Westermann, D.; Ravens, U. Altered physiological functions and ion currents in atrial fibroblasts from patients with chronic atrial fibrillation. Physiol. Rep. 2016, 4. [Google Scholar] [CrossRef] [PubMed]
  64. Rook, M.B.; van Ginneken, A.C.; de Jonge, B.; el Aoumari, A.; Gros, D.; Jongsma, H.J. Differences in gap junction channels between cardiac myocytes, fibroblasts, and heterologous pairs. Am. J. Physiol. Cell Physiol. 1992, 263, C959–C977. [Google Scholar] [CrossRef] [PubMed]
  65. Saito, T.; Fujiwara, Y.; Fujiwara, R.; Hasegawa, H.; Kibira, S.; Miura, H.; Miura, M. Role of augmented expression of intermediate-conductance Ca2+-activated K+ channels in postischaemic heart. Clin. Exp. Pharmacol. Physiol. 2002, 29, 324–329. [Google Scholar] [CrossRef] [PubMed]
  66. El Chemaly, A.; Guinamard, R.; Demion, M.; Fares, N.; Jebara, V.; Faivre, J.F.; Bois, P. A voltage-activated proton current in human cardiac fibroblasts. Biochem. Biophys. Res. Commun. 2006, 340, 512–516. [Google Scholar] [CrossRef] [PubMed]
  67. Mohis, M.; Edwards, S.; Ryan, S.; Rizvi, F.; Tajik, A.J.; Jahangir, A.; Ross, G.R. Aging-related increase in store-operated Ca(2+) influx in human ventricular fibroblasts. Am. J. Physiol. Heart Circ. Physiol. 2018, 315, H83–H91. [Google Scholar] [CrossRef] [PubMed]
  68. Ross, G.R.; Bajwa, T., Jr.; Edwards, S.; Emelyanova, L.; Rizvi, F.; Holmuhamedov, E.L.; Werner, P.; Downey, F.X.; Tajik, A.J.; Jahangir, A. Enhanced store-operated Ca(2+) influx and ORAI1 expression in ventricular fibroblasts from human failing heart. Biol. Open 2017, 6, 326–332. [Google Scholar] [CrossRef] [PubMed]
  69. Rose, R.A.; Hatano, N.; Ohya, S.; Imaizumi, Y.; Giles, W.R. C-type natriuretic peptide activates a non-selective cation current in acutely isolated rat cardiac fibroblasts via natriuretic peptide C receptor-mediated signalling. J. Physiol. 2007, 580, 255–274. [Google Scholar] [CrossRef] [PubMed]
  70. Harada, M.; Luo, X.; Qi, X.Y.; Tadevosyan, A.; Maguy, A.; Ordog, B.; Ledoux, J.; Kato, T.; Naud, P.; Voigt, N.; et al. Transient receptor potential canonical-3 channel-dependent fibroblast regulation in atrial fibrillation. Circulation 2012, 126, 2051–2064. [Google Scholar] [CrossRef] [PubMed]
  71. Hatano, N.; Itoh, Y.; Muraki, K. Cardiac fibroblasts have functional TRPV4 activated by 4alpha-phorbol 12,13-didecanoate. Life Sci. 2009, 85, 808–814. [Google Scholar] [CrossRef] [PubMed]
  72. Runnels, L.W.; Yue, L.; Clapham, D.E. The TRPM7 channel is inactivated by PIP(2) hydrolysis. Nat. Cell Biol. 2002, 4, 329–336. [Google Scholar] [CrossRef] [PubMed]
  73. Oguri, G.; Nakajima, T.; Yamamoto, Y.; Takano, N.; Tanaka, T.; Kikuchi, H.; Morita, T.; Nakamura, F.; Yamasoba, T.; Komuro, I. Effects of methylglyoxal on human cardiac fibroblast: Roles of transient receptor potential ankyrin 1 (TRPA1) channels. Am. J. Physiol. Heart Circ. Physiol. 2014, 307, H1339–H1352. [Google Scholar] [CrossRef] [PubMed]
  74. Vasquez, C.; Benamer, N.; Morley, G.E. The cardiac fibroblast: Functional and electrophysiological considerations in healthy and diseased hearts. J. Cardiovasc. Pharmacol. 2011, 57, 380–388. [Google Scholar] [CrossRef] [PubMed]
  75. Ongstad, E.; Kohl, P. Fibroblast-myocyte coupling in the heart: Potential relevance for therapeutic interventions. J. Mol. Cell. Cardiol. 2016, 91, 238–246. [Google Scholar] [CrossRef] [PubMed]
  76. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [PubMed]
  77. Oh-hora, M.; Rao, A. Calcium signaling in lymphocytes. Curr. Opin. Immunol. 2008, 20, 250–258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Clapham, D.E. Calcium signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [PubMed]
  79. Nowycky, M.C.; Thomas, A.P. Intracellular calcium signaling. J. Cell Sci. 2002, 115, 3715–3716. [Google Scholar] [CrossRef] [Green Version]
  80. Lee, H.C.; Aarhus, R. A derivative of NADP mobilizes calcium stores insensitive to inositol trisphosphate and cyclic ADP-ribose. J. Biol. Chem. 1995, 270, 2152–2157. [Google Scholar] [CrossRef]
  81. Hohenegger, M.; Suko, J.; Gscheidlinger, R.; Drobny, H.; Zidar, A. Nicotinic acid-adenine dinucleotide phosphate activates the skeletal muscle ryanodine receptor. Biochem. J. 2002, 367, 423–431. [Google Scholar] [CrossRef]
  82. Calcraft, P.J.; Ruas, M.; Pan, Z.; Cheng, X.; Arredouani, A.; Hao, X.; Tang, J.; Rietdorf, K.; Teboul, L.; Chuang, K.T.; et al. NAADP mobilizes calcium from acidic organelles through two-pore channels. Nature 2009, 459, 596–600. [Google Scholar] [CrossRef] [Green Version]
  83. Brailoiu, E.; Churamani, D.; Cai, X.; Schrlau, M.G.; Brailoiu, G.C.; Gao, X.; Hooper, R.; Boulware, M.J.; Dun, N.J.; Marchant, J.S.; et al. Essential requirement for two-pore channel 1 in NAADP-mediated calcium signaling. J. Cell Biol. 2009, 186, 201–209. [Google Scholar] [CrossRef] [PubMed]
  84. Pandey, V.; Chuang, C.C.; Lewis, A.M.; Aley, P.K.; Brailoiu, E.; Dun, N.J.; Churchill, G.C.; Patel, S. Recruitment of NAADP-sensitive acidic Ca2+ stores by glutamate. Biochem. J. 2009, 422, 503–512. [Google Scholar] [CrossRef] [PubMed]
  85. Ruas, M.; Rietdorf, K.; Arredouani, A.; Davis, L.C.; Lloyd-Evans, E.; Koegel, H.; Funnell, T.M.; Morgan, A.J.; Ward, J.A.; Watanabe, K.; et al. Purified TPC isoforms form NAADP receptors with distinct roles for Ca(2+) signaling and endolysosomal trafficking. Curr. Biol. CB 2010, 20, 703–709. [Google Scholar] [CrossRef] [PubMed]
  86. Guse, A.H.; Wolf, I.M. Ca(2+) microdomains, NAADP and type 1 ryanodine receptor in cell activation. Biochimica et Biophysica Acta 2016, 1863, 1379–1384. [Google Scholar] [CrossRef] [PubMed]
  87. Kiseleva, I.; Kamkin, A.; Kohl, P.; Lab, M.J. Calcium and mechanically induced potentials in fibroblasts of rat atrium. Cardiovasc. Res. 1996, 32, 98–111. [Google Scholar] [CrossRef] [Green Version]
  88. Ding, Z.; Yuan, J.; Liang, Y.; Wu, J.; Gong, H.; Ye, Y.; Jiang, G.; Yin, P.; Li, Y.; Zhang, G.; et al. Ryanodine Receptor Type 2 Plays a Role in the Development of Cardiac Fibrosis under Mechanical Stretch Through TGFbeta-1. Int. Heart J. 2017, 58, 957–961. [Google Scholar] [CrossRef] [PubMed]
  89. Higashida, H.; Egorova, A.; Higashida, C.; Zhong, Z.G.; Yokoyama, S.; Noda, M.; Zhang, J.S. Sympathetic potentiation of cyclic ADP-ribose formation in rat cardiac myocytes. J. Biol. Chem. 1999, 274, 33348–33354. [Google Scholar] [CrossRef] [PubMed]
  90. Gul, R.; Park, D.R.; Shawl, A.I.; Im, S.Y.; Nam, T.S.; Lee, S.H.; Ko, J.K.; Jang, K.Y.; Kim, D.; Kim, U.H. Nicotinic Acid Adenine Dinucleotide Phosphate (NAADP) and Cyclic ADP-Ribose (cADPR) Mediate Ca2+ Signaling in Cardiac Hypertrophy Induced by beta-Adrenergic Stimulation. PLoS ONE 2016, 11, e0149125. [Google Scholar] [CrossRef]
  91. Macgregor, A.; Yamasaki, M.; Rakovic, S.; Sanders, L.; Parkesh, R.; Churchill, G.C.; Galione, A.; Terrar, D.A. NAADP controls cross-talk between distinct Ca2+ stores in the heart. J. Biol. Chem. 2007, 282, 15302–15311. [Google Scholar] [CrossRef]
  92. Macgregor, A.T.; Rakovic, S.; Galione, A.; Terrar, D.A. Dual effects of cyclic ADP-ribose on sarcoplasmic reticulum Ca2+ release and storage in cardiac myocytes isolated from guinea-pig and rat ventricle. Cell Calcium 2007, 41, 537–546. [Google Scholar] [CrossRef]
  93. Snead, A.N.; Insel, P.A. Defining the cellular repertoire of GPCRs identifies a profibrotic role for the most highly expressed receptor, protease-activated receptor 1, in cardiac fibroblasts. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2012, 26, 4540–4547. [Google Scholar] [CrossRef] [PubMed]
  94. Brilla, C.G.; Scheer, C.; Rupp, H. Angiotensin II and intracellular calcium of adult cardiac fibroblasts. J. Mol. Cell. Cardiol. 1998, 30, 1237–1246. [Google Scholar] [CrossRef] [PubMed]
  95. Meszaros, J.G.; Gonzalez, A.M.; Endo-Mochizuki, Y.; Villegas, S.; Villarreal, F.; Brunton, L.L. Identification of G protein-coupled signaling pathways in cardiac fibroblasts: Cross talk between G(q) and G(s). Am. J. Physiol. Cell Physiol. 2000, 278, C154–C162. [Google Scholar] [CrossRef] [PubMed]
  96. Sun, Y.; Ramires, F.J.; Zhou, G.; Ganjam, V.K.; Weber, K.T. Fibrous tissue and angiotensin II. J. Mol. Cell. Cardiol. 1997, 29, 2001–2012. [Google Scholar] [CrossRef] [PubMed]
  97. Weber, K.T.; Sun, Y.; Bhattacharya, S.K.; Ahokas, R.A.; Gerling, I.C. Myofibroblast-mediated mechanisms of pathological remodelling of the heart. Nat. Rev. Cardiol. 2013, 10, 15–26. [Google Scholar] [CrossRef] [PubMed]
  98. Bomb, R.; Heckle, M.R.; Sun, Y.; Mancarella, S.; Guntaka, R.V.; Gerling, I.C.; Weber, K.T. Myofibroblast secretome and its auto-/paracrine signaling. Expert Rev. Cardiovasc. Ther. 2016, 14, 591–598. [Google Scholar] [CrossRef] [PubMed]
  99. Munoz-Rodriguez, C.; Fernandez, S.; Osorio, J.M.; Olivares, F.; Anfossi, R.; Bolivar, S.; Humeres, C.; Boza, P.; Vivar, R.; Pardo-Jimenez, V.; et al. Expression and function of TLR4- induced B1R bradykinin receptor on cardiac fibroblasts. Toxicol. Appl. Pharmacol. 2018, 351, 46–56. [Google Scholar] [CrossRef] [PubMed]
  100. Catalan, M.; Aranguiz, P.; Boza, P.; Olmedo, I.; Humeres, C.; Vivar, R.; Anfossi, R.; Ayala, P.; Espinoza, C.; Lavandero, S.; et al. TGF-beta1 induced up-regulation of B1 kinin receptor promotes antifibrotic activity in rat cardiac myofibroblasts. Mol. Biol. Rep. 2019. [Google Scholar] [CrossRef] [PubMed]
  101. Lu, D.; Insel, P.A. Cellular mechanisms of tissue fibrosis. 6. Purinergic signaling and response in fibroblasts and tissue fibrosis. Am. J. Physiol. Cell Physiol. 2014, 306, C779–C788. [Google Scholar] [CrossRef] [PubMed]
  102. Vecchio, E.A.; White, P.J.; May, L.T. Targeting Adenosine Receptors for the Treatment of Cardiac Fibrosis. Front. Pharmacol. 2017, 8, 243. [Google Scholar] [CrossRef] [PubMed]
  103. Vecchio, E.A.; Chuo, C.H.; Baltos, J.A.; Ford, L.; Scammells, P.J.; Wang, B.H.; Christopoulos, A.; White, P.J.; May, L.T. The hybrid molecule, VCP746, is a potent adenosine A2B receptor agonist that stimulates anti-fibrotic signalling. Biochem. Pharmacol. 2016, 117, 46–56. [Google Scholar] [CrossRef] [PubMed]
  104. Chen, J.B.; Liu, W.J.; Che, H.; Liu, J.; Sun, H.Y.; Li, G.R. Adenosine-5'-triphosphate up-regulates proliferation of human cardiac fibroblasts. Br. J. Pharmacol. 2012, 166, 1140–1150. [Google Scholar] [CrossRef] [PubMed]
  105. Antonioli, L.; Blandizzi, C.; Pacher, P.; Hasko, G. The Purinergic System as a Pharmacological Target for the Treatment of Immune-Mediated Inflammatory Diseases. Pharmacol. Rev. 2019, 71, 345–382. [Google Scholar] [CrossRef] [PubMed]
  106. Talasila, A.; Germack, R.; Dickenson, J.M. Characterization of P2Y receptor subtypes functionally expressed on neonatal rat cardiac myofibroblasts. Br. J. Pharmacol. 2009, 158, 339–353. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, X.; Zhang, T.; Wu, J.; Yu, X.; Zheng, D.; Yang, F.; Li, T.; Wang, L.; Zhao, Y.; Dong, S.; et al. Calcium sensing receptor promotes cardiac fibroblast proliferation and extracellular matrix secretion. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2014, 33, 557–568. [Google Scholar] [CrossRef] [PubMed]
  108. Liao, Y.; Plummer, N.W.; George, M.D.; Abramowitz, J.; Zhu, M.X.; Birnbaumer, L. A role for Orai in TRPC-mediated Ca2+ entry suggests that a TRPC:Orai complex may mediate store and receptor operated Ca2+ entry. Proc. Natl. Acad. Sci. USA 2009, 106, 3202–3206. [Google Scholar] [CrossRef] [PubMed]
  109. Mukherjee, S.; Ayaub, E.A.; Murphy, J.; Lu, C.; Kolb, M.; Ask, K.; Janssen, L.J. Disruption of Calcium Signaling in Fibroblasts and Attenuation of Bleomycin-Induced Fibrosis by Nifedipine. Am. J. Respir. Cell Mol. Biol. 2015, 53, 450–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Surprenant, A.; North, R.A. Signaling at purinergic P2X receptors. Annu. Rev. Physiol. 2009, 71, 333–359. [Google Scholar] [CrossRef] [PubMed]
  111. Solini, A.; Chiozzi, P.; Morelli, A.; Fellin, R.; Di Virgilio, F. Human primary fibroblasts in vitro express a purinergic P2X7 receptor coupled to ion fluxes, microvesicle formation and IL-6 release. J. Cell Sci. 1999, 112 Pt 3, 297–305. [Google Scholar]
  112. Gentile, D.; Natale, M.; Lazzerini, P.E.; Capecchi, P.L.; Laghi-Pasini, F. The role of P2X7 receptors in tissue fibrosis: A brief review. Purinergic Signal. 2015, 11, 435–440. [Google Scholar] [CrossRef] [PubMed]
  113. Therkildsen, J.R.; Christensen, M.G.; Tingskov, S.J.; Wehmoller, J.; Norregaard, R.; Praetorius, H.A. Lack of P2X7 Receptors Protects against Renal Fibrosis after Pyelonephritis with alpha-Hemolysin-Producing Escherichia coli. Am. J. Pathol. 2019, 189, 1201–1211. [Google Scholar] [CrossRef] [PubMed]
  114. Yang, R.; Liang, B.T. Cardiac P2X(4) receptors: Targets in ischemia and heart failure? Circ. Res. 2012, 111, 397–401. [Google Scholar] [CrossRef] [PubMed]
  115. Yang, T.; Shen, J.B.; Yang, R.; Redden, J.; Dodge-Kafka, K.; Grady, J.; Jacobson, K.A.; Liang, B.T. Novel protective role of endogenous cardiac myocyte P2X4 receptors in heart failure. Circ. Heart Fail. 2014, 7, 510–518. [Google Scholar] [CrossRef] [PubMed]
  116. Novitskaya, T.; Chepurko, E.; Covarrubias, R.; Novitskiy, S.; Ryzhov, S.V.; Feoktistov, I.; Gumina, R.J. Extracellular nucleotide regulation and signaling in cardiac fibrosis. J. Mol. Cell. Cardiol. 2016, 93, 47–56. [Google Scholar] [CrossRef] [PubMed]
  117. Lewis, R.S. The molecular choreography of a store-operated calcium channel. Nature 2007, 446, 284–287. [Google Scholar] [CrossRef] [PubMed]
  118. Prakriya, M.; Lewis, R.S. Store-operated calcium channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef]
  119. Feske, S. CRAC channels and disease—From human CRAC channelopathies and animal models to novel drugs. Cell Calcium 2019, 80, 112–116. [Google Scholar] [CrossRef] [PubMed]
  120. Rosenberg, P.; Katz, D.; Bryson, V. SOCE and STIM1 signaling in the heart: Timing and location matter. Cell Calcium 2019, 77, 20–28. [Google Scholar] [CrossRef]
  121. Voelkers, M.; Salz, M.; Herzog, N.; Frank, D.; Dolatabadi, N.; Frey, N.; Gude, N.; Friedrich, O.; Koch, W.J.; Katus, H.A.; et al. Orai1 and Stim1 regulate normal and hypertrophic growth in cardiomyocytes. J. Mol. Cell. Cardiol. 2010, 48, 1329–1334. [Google Scholar] [CrossRef] [Green Version]
  122. Correll, R.N.; Goonasekera, S.A.; van Berlo, J.H.; Burr, A.R.; Accornero, F.; Zhang, H.; Makarewich, C.A.; York, A.J.; Sargent, M.A.; Chen, X.; et al. STIM1 elevation in the heart results in aberrant Ca(2)(+) handling and cardiomyopathy. J. Mol. Cell. Cardiol. 2015, 87, 38–47. [Google Scholar] [CrossRef]
  123. Horton, J.S.; Buckley, C.L.; Alvarez, E.M.; Schorlemmer, A.; Stokes, A.J. The calcium release-activated calcium channel Orai1 represents a crucial component in hypertrophic compensation and the development of dilated cardiomyopathy. Channels 2014, 8, 35–48. [Google Scholar] [CrossRef] [PubMed]
  124. Bartoli, F.; Sabourin, J. Cardiac Remodeling and Disease: Current Understanding of STIM1/Orai1-Mediated Store-Operated Ca(2+) Entry in Cardiac Function and Pathology. Adv. Exp. Med. Biol. 2017, 993, 523–534. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, B.; Jiang, J.; Yue, Z.; Liu, S.; Ma, Y.; Yu, N.; Gao, Y.; Sun, S.; Chen, S.; Liu, P. Store-Operated Ca(2+) Entry (SOCE) contributes to angiotensin II-induced cardiac fibrosis in cardiac fibroblasts. J. Pharmacol. Sci. 2016, 132, 171–180. [Google Scholar] [CrossRef] [PubMed]
  126. Montell, C.; Birnbaumer, L.; Flockerzi, V. The TRP channels, a remarkably functional family. Cell 2002, 108, 595–598. [Google Scholar] [CrossRef]
  127. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524. [Google Scholar] [CrossRef] [PubMed]
  128. Nilius, B.; Voets, T.; Peters, J. TRP Channels in Disease. Sci. STKE 2005, 2005, re8. [Google Scholar] [CrossRef] [PubMed]
  129. Story, G.M.; Peier, A.M.; Reeve, A.J.; Eid, S.R.; Mosbacher, J.; Hricik, T.R.; Earley, T.J.; Hergarden, A.C.; Andersson, D.A.; Hwang, S.W.; et al. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 2003, 112, 819–829. [Google Scholar] [CrossRef]
  130. Jordt, S.E.; Bautista, D.M.; Chuang, H.H.; McKemy, D.D.; Zygmunt, P.M.; Hogestatt, E.D.; Meng, I.D.; Julius, D. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature 2004, 427, 260–265. [Google Scholar] [CrossRef] [PubMed]
  131. Tominaga, M.; Caterina, M.J. Thermosensation and pain. J. Neurobiol. 2004, 61, 3–12. [Google Scholar] [CrossRef]
  132. Chubanov, V.; Mederos, Y.S.M.; Waring, J.; Plank, A.; Gudermann, T. Emerging roles of TRPM6/TRPM7 channel kinase signal transduction complexes. Naunyn-Schmiedeberg's Arch. Pharmacol. 2005, 371, 334–341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Dong, X.P.; Wang, X.; Xu, H. TRP channels of intracellular membranes. J. Neurochem. 2010, 113, 313–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Yue, L.; Peng, J.B.; Hediger, M.A.; Clapham, D.E. CaT1 manifests the pore properties of the calcium-release-activated calcium channel. Nature 2001, 410, 705–709. [Google Scholar] [CrossRef] [PubMed]
  135. Nilius, B.; Vennekens, R.; Prenen, J.; Hoenderop, J.G.; Droogmans, G.; Bindels, R.J. The single pore residue Asp542 determines Ca2+ permeation and Mg2+ block of the epithelial Ca2+ channel. J. Biol. Chem. 2001, 276, 1020–1025. [Google Scholar] [CrossRef] [PubMed]
  136. 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]
  137. Li, M.; Jiang, J.; Yue, L. Functional Characterization of Homo- and Heteromeric Channel Kinases TRPM6 and TRPM7. J. Gen. Physiol. 2006, 127, 525–537. [Google Scholar] [CrossRef] [Green Version]
  138. Schlingmann, K.P.; Waldegger, S.; Konrad, M.; Chubanov, V.; Gudermann, T. TRPM6 and TRPM7--Gatekeepers of human magnesium metabolism. Biochimica Et Biophysica Acta 2007, 1772, 813–821. [Google Scholar] [CrossRef]
  139. Venkatachalam, K.; Montell, C. TRP Channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef] [Green Version]
  140. Zimmermann, K.; Lennerz, J.K.; Hein, A.; Link, A.S.; Kaczmarek, J.S.; Delling, M.; Uysal, S.; Pfeifer, J.D.; Riccio, A.; Clapham, D.E. Transient receptor potential cation channel, subfamily C, member 5 (TRPC5) is a cold-transducer in the peripheral nervous system. Proc. Natl. Acad. Sci. USA 2011, 108, 18114–18119. [Google Scholar] [CrossRef] [Green Version]
  141. Nilius, B. TRP channels in disease. Biochimica Et Biophysica Acta 2007, 1772, 805–812. [Google Scholar] [CrossRef] [Green Version]
  142. Vriens, J.; Owsianik, G.; Hofmann, T.; Philipp, S.E.; Stab, J.; Chen, X.; Benoit, M.; Xue, F.; Janssens, A.; Kerselaers, S.; et al. TRPM3 is a nociceptor channel involved in the detection of noxious heat. Neuron 2011, 70, 482–494. [Google Scholar] [CrossRef]
  143. Zufall, F.; Ukhanov, K.; Lucas, P.; Liman, E.R.; Leinders-Zufall, T. Neurobiology of TRPC2: From gene to behavior. Pflug. Arch. Eur. J. Physiol. 2005, 451, 61–71. [Google Scholar] [CrossRef] [PubMed]
  144. Hofmann, T.; Obukhov, A.G.; Schaefer, M.; Harteneck, C.; Gudermann, T.; Schultz, G. Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature 1999, 397, 259–263. [Google Scholar] [CrossRef] [PubMed]
  145. Beck, B.; Zholos, A.; Sydorenko, V.; Roudbaraki, M.; Lehen’kyi, V.; Bordat, P.; Prevarskaya, N.; Skryma, R. TRPC7 is a receptor-operated DAG-activated channel in human keratinocytes. J. Investig. Dermatol. 2006, 126, 1982–1993. [Google Scholar] [CrossRef] [PubMed]
  146. Vangeel, L.; Voets, T. Transient Receptor Potential Channels and Calcium Signaling. Cold Spring Harb. Perspect. Biol. 2019, 11. [Google Scholar] [CrossRef] [PubMed]
  147. Hofmann, T.; Schaefer, M.; Schultz, G.; Gudermann, T. Subunit composition of mammalian transient receptor potential channels in living cells. Proc. Natl. Acad. Sci. USA 2002, 99, 7461–7466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Nadler, M.J.; Hermosura, M.C.; Inabe, K.; Perraud, A.L.; Zhu, Q.; Stokes, A.J.; Kurosaki, T.; Kinet, J.P.; Penner, R.; Scharenberg, A.M.; et al. LTRPC7 is a Mg.ATP-regulated divalent cation channel required for cell viability. Nature 2001, 411, 590–595. [Google Scholar] [CrossRef] [PubMed]
  149. Li, M.; Du, J.; Jiang, J.; Ratzan, W.; Su, L.-T.; Runnels, L.W.; Yue, L. Molecular Determinants of Mg2+ and Ca2+ Permeability and pH Sensitivity in TRPM6 and TRPM7. J. Biol. Chem. 2007, 282, 25817–25830. [Google Scholar] [CrossRef]
  150. Voets, T.; Nilius, B.; Hoefs, S.; van der Kemp, A.W.C.M.; Droogmans, G.; Bindels, R.J.M.; Hoenderop, J.G.J. TRPM6 Forms the Mg2+ Influx Channel Involved in Intestinal and Renal Mg2+ Absorption. J. Biol. Chem. 2004, 279, 19–25. [Google Scholar] [CrossRef] [Green Version]
  151. 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]
  152. Prawitt, D.; Monteilh-Zoller, M.K.; Brixel, L.; Spangenberg, C.; Zabel, B.; Fleig, A.; Penner, R. TRPM5 is a transient Ca2+-activated cation channel responding to rapid changes in [Ca2+]i. Proc. Natl. Acad. Sci. USA 2003, 100, 15166–15171. [Google Scholar] [CrossRef]
  153. Liu, D.; Liman, E.R. Intracellular Ca2+ and the phospholipid PIP2 regulate the taste transduction ion channel TRPM5. Proc. Natl. Acad. Sci. USA 2003, 100, 15160–15165. [Google Scholar] [CrossRef] [PubMed]
  154. Du, J.; Xie, J.; Yue, L. Intracellular calcium activates TRPM2 and its alternative spliced isoforms. Proc. Natl. Acad. Sci. USA 2009, 107, 7239–7244. [Google Scholar] [CrossRef] [PubMed]
  155. Zurborg, S.; Yurgionas, B.; Jira, J.A.; Caspani, O.; Heppenstall, P.A. Direct activation of the ion channel TRPA1 by Ca2+. Nat. Neurosci. 2007, 10, 277–279. [Google Scholar] [CrossRef] [PubMed]
  156. Hara, Y.; Wakamori, M.; Ishii, M.; Maeno, E.; Nishida, M.; Yoshida, T.; Yamada, H.; Shimizu, S.; Mori, E.; Kudoh, J.; et al. LTRPC2 Ca2+-permeable channel activated by changes in redox status confers susceptibility to cell death. Mol. Cell 2002, 9, 163–173. [Google Scholar] [CrossRef]
  157. Perraud, A.L.; Fleig, A.; Dunn, C.A.; Bagley, L.A.; Launay, P.; Schmitz, C.; Stokes, A.J.; Zhu, Q.; Bessman, M.J.; Penner, R.; et al. ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature 2001, 411, 595–599. [Google Scholar] [CrossRef]
  158. Sano, Y.; Inamura, K.; Miyake, A.; Mochizuki, S.; Yokoi, H.; Matsushime, H.; Furuichi, K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science 2001, 293, 1327–1330. [Google Scholar] [CrossRef] [PubMed]
  159. Yue, Z.; Xie, J.; Yu, A.S.; Stock, J.; Du, J.; Yue, L. Role of TRP channels in the cardiovascular system. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H157–H182. [Google Scholar] [CrossRef]
  160. Nishida, M.; Onohara, N.; Sato, Y.; Suda, R.; Ogushi, M.; Tanabe, S.; Inoue, R.; Mori, Y.; Kurose, H. Galpha12/13-mediated up-regulation of TRPC6 negatively regulates endothelin-1-induced cardiac myofibroblast formation and collagen synthesis through nuclear factor of activated T cells activation. J. Biol. Chem. 2007, 282, 23117–23128. [Google Scholar] [CrossRef]
  161. Xie, J.; Sun, B.; Du, J.; Yang, W.; Chen, H.-C.; Overton, J.D.; Runnels, L.W.; Yue, L. Phosphatidylinositol 4, 5-Bisphosphate (PIP2) Controls Magnesium Gatekeeper TRPM6 Activity. Biophys. J. 2012, 102, 24a. [Google Scholar] [CrossRef]
  162. Chen, H.C.; Xie, J.; Zhang, Z.; Su, L.T.; Yue, L.; Runnels, L.W. Blockade of TRPM7 channel activity and cell death by inhibitors of 5-lipoxygenase. PLoS ONE 2010, 5, e11161. [Google Scholar] [CrossRef]
  163. Falcon, D.; Galeano-Otero, I.; Calderon-Sanchez, E.; Del Toro, R.; Martin-Bornez, M.; Rosado, J.A.; Hmadcha, A.; Smani, T. TRP Channels: Current Perspectives in the Adverse Cardiac Remodeling. Front. Physiol. 2019, 10, 159. [Google Scholar] [CrossRef] [PubMed]
  164. Davis, J.; Burr, A.R.; Davis, G.F.; Birnbaumer, L.; Molkentin, J.D. A TRPC6-dependent pathway for myofibroblast transdifferentiation and wound healing in vivo. Dev. Cell 2012, 23, 705–715. [Google Scholar] [CrossRef] [PubMed]
  165. Freichel, M.; Berlin, M.; Schurger, A.; Mathar, I.; Bacmeister, L.; Medert, R.; Frede, W.; Marx, A.; Segin, S.; Londono, J.E.C. TRP Channels in the Heart. In Neurobiology of TRP Channels, 2nd ed.; Emir, T.L.R., Ed.; CRC Press: Boca Raton, FL, USA, 2017; pp. 149–185. [Google Scholar]
  166. Kiyonaka, S.; Kato, K.; Nishida, M.; Mio, K.; Numaga, T.; Sawaguchi, Y.; Yoshida, T.; Wakamori, M.; Mori, E.; Numata, T.; et al. Selective and direct inhibition of TRPC3 channels underlies biological activities of a pyrazole compound. Proc. Natl. Acad. Sci. USA 2009, 106, 5400–5405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Numaga-Tomita, T.; Oda, S.; Shimauchi, T.; Nishimura, A.; Mangmool, S.; Nishida, M. TRPC3 Channels in Cardiac Fibrosis. Front. Cardiovasc. Med. 2017, 4, 56. [Google Scholar] [CrossRef] [PubMed]
  168. Kitajima, N.; Numaga-Tomita, T.; Watanabe, M.; Kuroda, T.; Nishimura, A.; Miyano, K.; Yasuda, S.; Kuwahara, K.; Sato, Y.; Ide, T.; et al. TRPC3 positively regulates reactive oxygen species driving maladaptive cardiac remodeling. Sci. Rep. 2016, 6, 37001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Numaga-Tomita, T.; Kitajima, N.; Kuroda, T.; Nishimura, A.; Miyano, K.; Yasuda, S.; Kuwahara, K.; Sato, Y.; Ide, T.; Birnbaumer, L.; et al. TRPC3-GEF-H1 axis mediates pressure overload-induced cardiac fibrosis. Sci. Rep. 2016, 6, 39383. [Google Scholar] [CrossRef] [PubMed]
  170. Saliba, Y.; Karam, R.; Smayra, V.; Aftimos, G.; Abramowitz, J.; Birnbaumer, L.; Fares, N. Evidence of a Role for Fibroblast Transient Receptor Potential Canonical 3 Ca2+ Channel in Renal Fibrosis. J. Am. Soc. Nephrol. 2015, 26, 1855–1876. [Google Scholar] [CrossRef]
  171. Ikeda, K.; Nakajima, T.; Yamamoto, Y.; Takano, N.; Tanaka, T.; Kikuchi, H.; Oguri, G.; Morita, T.; Nakamura, F.; Komuro, I. Roles of transient receptor potential canonical (TRPC) channels and reverse-mode Na+/Ca2+ exchanger on cell proliferation in human cardiac fibroblasts: Effects of transforming growth factor beta1. Cell Calcium 2013, 54, 213–225. [Google Scholar] [CrossRef]
  172. Kapur, N.K.; Qiao, X.; Paruchuri, V.; Mackey, E.E.; Daly, G.H.; Ughreja, K.; Morine, K.J.; Levine, J.; Aronovitz, M.J.; Hill, N.S.; et al. Reducing endoglin activity limits calcineurin and TRPC-6 expression and improves survival in a mouse model of right ventricular pressure overload. J. Am. Heart Assoc. 2014, 3. [Google Scholar] [CrossRef]
  173. Wang, Q.; Ma, S.; Li, D.; Zhang, Y.; Tang, B.; Qiu, C.; Yang, Y.; Yang, D. Dietary capsaicin ameliorates pressure overload-induced cardiac hypertrophy and fibrosis through the transient receptor potential vanilloid type 1. Am. J. Hypertens. 2014, 27, 1521–1529. [Google Scholar] [CrossRef]
  174. Wang, Q.; Zhang, Y.; Li, D.; Zhang, Y.; Tang, B.; Li, G.; Yang, Y.; Yang, D. Transgenic overexpression of transient receptor potential vanilloid subtype 1 attenuates isoproterenol-induced myocardial fibrosis in mice. Int. J. Mol. Med. 2016, 38, 601–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Sexton, A.; McDonald, M.; Cayla, C.; Thiemermann, C.; Ahluwalia, A. 12-Lipoxygenase-derived eicosanoids protect against myocardial ischemia/reperfusion injury via activation of neuronal TRPV1. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2007, 21, 2695–2703. [Google Scholar] [CrossRef]
  176. Wang, L.; Wang, D.H. TRPV1 gene knockout impairs postischemic recovery in isolated perfused heart in mice. Circulation 2005, 112, 3617–3623. [Google Scholar] [CrossRef] [PubMed]
  177. Zhong, B.; Wang, D.H. TRPV1 gene knockout impairs preconditioning protection against myocardial injury in isolated perfused hearts in mice. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H1791–H1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Iwata, Y.; Ohtake, H.; Suzuki, O.; Matsuda, J.; Komamura, K.; Wakabayashi, S. Blockade of sarcolemmal TRPV2 accumulation inhibits progression of dilated cardiomyopathy. Cardiovasc. Res. 2013, 99, 760–768. [Google Scholar] [CrossRef]
  179. Koch, S.E.; Mann, A.; Jones, S.; Robbins, N.; Alkhattabi, A.; Worley, M.C.; Gao, X.; Lasko-Roiniotis, V.M.; Karani, R.; Fulford, L.; et al. Transient receptor potential vanilloid 2 function regulates cardiac hypertrophy via stretch-induced activation. J. Hypertens. 2017, 35, 602–611. [Google Scholar] [CrossRef] [PubMed]
  180. Koch, S.E.; Nieman, M.L.; Robbins, N.; Slone, S.; Worley, M.; Green, L.C.; Chen, Y.; Barlow, A.; Tranter, M.; Wang, H.; et al. Tranilast Blunts the Hypertrophic and Fibrotic Response to Increased Afterload Independent of Cardiomyocyte Transient Receptor Potential Vanilloid 2 Channels. J. Cardiovasc. Pharmacol. 2018, 72, 40–48. [Google Scholar] [CrossRef]
  181. Entin-Meer, M.; Cohen, L.; Hertzberg-Bigelman, E.; Levy, R.; Ben-Shoshan, J.; Keren, G. TRPV2 knockout mice demonstrate an improved cardiac performance following myocardial infarction due to attenuated activity of peri-infarct macrophages. PLoS ONE 2017, 12, e0177132. [Google Scholar] [CrossRef]
  182. Zhang, Q.; Qi, H.; Cao, Y.; Shi, P.; Song, C.; Ba, L.; Chen, Y.; Gao, J.; Li, S.; Li, B.; et al. Activation of transient receptor potential vanilloid 3 channel (TRPV3) aggravated pathological cardiac hypertrophy via calcineurin/NFATc3 pathway in rats. J. Cell. Mol. Med. 2018, 22, 6055–6067. [Google Scholar] [CrossRef] [Green Version]
  183. Liu, Y.; Qi, H.; E, M.; Shi, P.; Zhang, Q.; Li, S.; Wang, Y.; Cao, Y.; Chen, Y.; Ba, L.; et al. Transient receptor potential vanilloid-3 (TRPV3) activation plays a central role in cardiac fibrosis induced by pressure overload in rats via TGF-beta1 pathway. Naunyn-Schmiedeberg's Arch. Pharmacol. 2018, 391, 131–143. [Google Scholar] [CrossRef]
  184. Thodeti, C.K.; Paruchuri, S.; Meszaros, J.G. A TRP to cardiac fibroblast differentiation. Channels 2013, 7, 211–214. [Google Scholar] [CrossRef] [Green Version]
  185. Adapala, R.K.; Minasyan, A.; Kanugula, A.K.; Cappelli, H.C.; Paruchuri, S.; Meszaros, G.J.; Thodeti, C.K. Targeting Trpv4 Channels Protects Heart From Pathological Remodeling Following Myocardial Infarction. Circulation 2017, 136 (Suppl. 1), A24061. [Google Scholar]
  186. Miller, B.A.; Wang, J.; Hirschler-Laszkiewicz, I.; Gao, E.; Song, J.; Zhang, X.Q.; Koch, W.J.; Madesh, M.; Mallilankaraman, K.; Gu, T.; et al. The second member of transient receptor potential-melastatin channel family protects hearts from ischemia-reperfusion injury. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H1010–H1022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Hoffman, N.E.; Miller, B.A.; Wang, J.; Elrod, J.W.; Rajan, S.; Gao, E.; Song, J.; Zhang, X.Q.; Hirschler-Laszkiewicz, I.; Shanmughapriya, S.; et al. Ca(2)(+) entry via Trpm2 is essential for cardiac myocyte bioenergetics maintenance. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H637–H650. [Google Scholar] [CrossRef] [PubMed]
  188. Yang, K.T.; Chang, W.L.; Yang, P.C.; Chien, C.L.; Lai, M.S.; Su, M.J.; Wu, M.L. Activation of the transient receptor potential M2 channel and poly(ADP-ribose) polymerase is involved in oxidative stress-induced cardiomyocyte death. Cell Death Differ. 2006, 13, 1815–1826. [Google Scholar] [CrossRef] [PubMed]
  189. Hiroi, T.; Wajima, T.; Negoro, T.; Ishii, M.; Nakano, Y.; Kiuchi, Y.; Mori, Y.; Shimizu, S. Neutrophil TRPM2 channels are implicated in the exacerbation of myocardial ischaemia/reperfusion injury. Cardiovasc. Res. 2013, 97, 271–281. [Google Scholar] [CrossRef] [PubMed]
  190. Sah, R.; Mesirca, P.; Mason, X.; Gibson, W.; Bates-Withers, C.; Van den Boogert, M.; Chaudhuri, D.; Pu, W.T.; Mangoni, M.E.; Clapham, D.E. Timing of myocardial trpm7 deletion during cardiogenesis variably disrupts adult ventricular function, conduction, and repolarization. Circulation 2013, 128, 101–114. [Google Scholar] [CrossRef]
  191. Sah, R.; Mesirca, P.; Van den Boogert, M.; Rosen, J.; Mably, J.; Mangoni, M.E.; Clapham, D.E. Ion channel-kinase TRPM7 is required for maintaining cardiac automaticity. Proc. Natl. Acad. Sci. USA 2013, 110, E3037–E3046. [Google Scholar] [CrossRef] [PubMed]
  192. Zhong, H.; Wang, T.; Lian, G.; Xu, C.; Wang, H.; Xie, L. TRPM7 regulates angiotensin II-induced sinoatrial node fibrosis in sick sinus syndrome rats by mediating Smad signaling. Heart Vessel. 2018, 33, 1094–1105. [Google Scholar] [CrossRef] [Green Version]
  193. Zhou, Y.; Yi, X.; Wang, T.; Li, M. Effects of angiotensin II on transient receptor potential melastatin 7 channel function in cardiac fibroblasts. Exp. Ther. Med. 2015, 9, 2008–2012. [Google Scholar] [CrossRef]
  194. Li, S.; Li, M.; Yi, X.; Guo, F.; Zhou, Y.; Chen, S.; Wu, X. TRPM7 channels mediate the functional changes in cardiac fibroblasts induced by angiotensin II. Int. J. Mol. Med. 2017, 39, 1291–1298. [Google Scholar] [CrossRef]
  195. Guo, J.L.; Yu, Y.; Jia, Y.Y.; Ma, Y.Z.; Zhang, B.Y.; Liu, P.Q.; Chen, S.R.; Jiang, J.M. Transient receptor potential melastatin 7 (TRPM7) contributes to H2O2-induced cardiac fibrosis via mediating Ca(2+) influx and extracellular signal-regulated kinase 1/2 (ERK1/2) activation in cardiac fibroblasts. J. Pharmacol. Sci. 2014, 125, 184–192. [Google Scholar] [CrossRef]
  196. Yu, Y.; Chen, S.; Xiao, C.; Jia, Y.; Guo, J.; Jiang, J.; Liu, P. TRPM7 is involved in angiotensin II induced cardiac fibrosis development by mediating calcium and magnesium influx. Cell Calcium 2014, 55, 252–260. [Google Scholar] [CrossRef]
  197. Andrei, S.R.; Ghosh, M.; Sinharoy, P.; Dey, S.; Bratz, I.N.; Damron, D.S. TRPA1 ion channel stimulation enhances cardiomyocyte contractile function via a CaMKII-dependent pathway. Channels 2017, 11, 587–603. [Google Scholar] [CrossRef]
  198. Andrei, S.R.; Sinharoy, P.; Bratz, I.N.; Damron, D.S. TRPA1 is functionally co-expressed with TRPV1 in cardiac muscle: Co-localization at z-discs, costameres and intercalated discs. Channels 2016, 10, 395–409. [Google Scholar] [CrossRef]
  199. Li, S.; Sun, X.; Wu, H.; Yu, P.; Wang, X.; Jiang, Z.; Gao, E.; Chen, J.; Li, D.; Qiu, C.; et al. TRPA1 Promotes Cardiac Myofibroblast Transdifferentiation after Myocardial Infarction Injury via the Calcineurin-NFAT-DYRK1A Signaling Pathway. Oxidative Med. Cell. Longev. 2019, 2019, 6408352. [Google Scholar] [CrossRef]
  200. Wang, Z.; Xu, Y.; Wang, M.; Ye, J.; Liu, J.; Jiang, H.; Ye, D.; Wan, J. TRPA1 inhibition ameliorates pressure overload-induced cardiac hypertrophy and fibrosis in mice. EBioMedicine 2018, 36, 54–62. [Google Scholar] [CrossRef] [Green Version]
  201. Eder, P.; Molkentin, J.D. TRPC channels as effectors of cardiac hypertrophy. Circ. Res. 2011, 108, 265–272. [Google Scholar] [CrossRef]
  202. Camacho Londono, J.E.; Tian, Q.; Hammer, K.; Schroder, L.; Camacho Londono, J.; Reil, J.C.; He, T.; Oberhofer, M.; Mannebach, S.; Mathar, I.; et al. A background Ca2+ entry pathway mediated by TRPC1/TRPC4 is critical for development of pathological cardiac remodelling. Eur. Heart J. 2015, 36, 2257–2266. [Google Scholar] [CrossRef]
  203. Wu, X.; Eder, P.; Chang, B.; Molkentin, J.D. TRPC channels are necessary mediators of pathologic cardiac hypertrophy. Proc. Natl. Acad. Sci. USA 2010, 107, 7000–7005. [Google Scholar] [CrossRef] [Green Version]
  204. Seo, K.; Rainer, P.P.; Shalkey Hahn, V.; Lee, D.I.; Jo, S.H.; Andersen, A.; Liu, T.; Xu, X.; Willette, R.N.; Lepore, J.J.; et al. Combined TRPC3 and TRPC6 blockade by selective small-molecule or genetic deletion inhibits pathological cardiac hypertrophy. Proc. Natl. Acad. Sci. USA 2014, 111, 1551–1556. [Google Scholar] [CrossRef] [Green Version]
  205. Certal, M.; Vinhas, A.; Barros-Barbosa, A.; Ferreirinha, F.; Costa, M.A.; Correia-de-Sa, P. ADP-Induced Ca(2+) Signaling and Proliferation of Rat Ventricular Myofibroblasts Depend on Phospholipase C-Linked TRP Channels Activation Within Lipid Rafts. J. Cell. Physiol. 2017, 232, 1511–1526. [Google Scholar] [CrossRef]
  206. Alfulaij, N.; Meiners, F.; Michalek, J.; Small-Howard, A.L.; Turner, H.C.; Stokes, A.J. Cannabinoids, the Heart of the Matter. J. Am. Heart Assoc. 2018, 7. [Google Scholar] [CrossRef]
  207. Huang, W.; Rubinstein, J.; Prieto, A.R.; Wang, D.H. Enhanced postmyocardial infarction fibrosis via stimulation of the transforming growth factor-beta-Smad2 signaling pathway: Role of transient receptor potential vanilloid type 1 channels. J. Hypertens. 2010, 28, 367–376. [Google Scholar] [CrossRef]
  208. Horton, J.S.; Buckley, C.L.; Stokes, A.J. Successful TRPV1 antagonist treatment for cardiac hypertrophy and heart failure in mice. Channels 2013, 7, 17–22. [Google Scholar] [CrossRef] [Green Version]
  209. Buckley, C.L.; Stokes, A.J. Mice lacking functional TRPV1 are protected from pressure overload cardiac hypertrophy. Channels 2011, 5, 367–374. [Google Scholar] [CrossRef] [Green Version]
  210. Jones, S.; Mann, A.; Worley, M.C.; Fulford, L.; Hall, D.; Karani, R.; Jiang, M.; Robbins, N.; Rubinstein, J.; Koch, S.E. The role of transient receptor potential vanilloid 2 channel in cardiac aging. Aging Clin. Exp. Res. 2017, 29, 863–873. [Google Scholar] [CrossRef]
  211. Ishii, T.; Uchida, K.; Hata, S.; Hatta, M.; Kita, T.; Miyake, Y.; Okamura, K.; Tamaoki, S.; Ishikawa, H.; Yamazaki, J. TRPV2 channel inhibitors attenuate fibroblast differentiation and contraction mediated by keratinocyte-derived TGF-beta1 in an in vitro wound healing model of rats. J. Dermatol. Sci. 2018, 90, 332–342. [Google Scholar] [CrossRef]
  212. Rahaman, S.O.; Grove, L.M.; Paruchuri, S.; Southern, B.D.; Abraham, S.; Niese, K.A.; Scheraga, R.G.; Ghosh, S.; Thodeti, C.K.; Zhang, D.X.; et al. TRPV4 mediates myofibroblast differentiation and pulmonary fibrosis in mice. J. Clin. Investig. 2014, 124, 5225–5238. [Google Scholar] [CrossRef]
  213. Kruse, M.; Schulze-Bahr, E.; Corfield, V.; Beckmann, A.; Stallmeyer, B.; Kurtbay, G.; Ohmert, I.; Brink, P.; Pongs, O. Impaired endocytosis of the ion channel TRPM4 is associated with human progressive familial heart block type I. J. Clin. Investig. 2009, 119, 2737–2744. [Google Scholar] [CrossRef]
  214. Liu, H.; Chatel, S.; Simard, C.; Syam, N.; Salle, L.; Probst, V.; Morel, J.; Millat, G.; Lopez, M.; Abriel, H.; et al. Molecular genetics and functional anomalies in a series of 248 Brugada cases with 11 mutations in the TRPM4 channel. PLoS ONE 2013, 8, e54131. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, H.; El Zein, L.; Kruse, M.; Guinamard, R.; Beckmann, A.; Bozio, A.; Kurtbay, G.; Megarbane, A.; Ohmert, I.; Blaysat, G.; et al. Gain-of-function mutations in TRPM4 cause autosomal dominant isolated cardiac conduction disease. Circ. Cardiovasc. Genet. 2010, 3, 374–385. [Google Scholar] [CrossRef] [PubMed]
  216. Stallmeyer, B.; Zumhagen, S.; Denjoy, I.; Duthoit, G.; Hebert, J.L.; Ferrer, X.; Maugenre, S.; Schmitz, W.; Kirchhefer, U.; Schulze-Bahr, E.; et al. Mutational spectrum in the Ca(2+)--activated cation channel gene TRPM4 in patients with cardiac conductance disturbances. Hum. Mutat. 2012, 33, 109–117. [Google Scholar] [CrossRef] [PubMed]
  217. Jin, J.; Desai, B.N.; Navarro, B.; Donovan, A.; Andrews, N.C.; Clapham, D.E. Deletion of Trpm7 disrupts embryonic development and thymopoiesis without altering Mg2+ homeostasis. Science 2008, 322, 756–760. [Google Scholar] [CrossRef]
  218. Jin, J.; Wu, L.J.; Jun, J.; Cheng, X.; Xu, H.; Andrews, N.C.; Clapham, D.E. The channel kinase, TRPM7, is required for early embryonic development. Proc. Natl. Acad. Sci. USA 2012, 109, E225–E233. [Google Scholar] [CrossRef] [PubMed]
  219. Rios, F.J.; Zou, Z.G.; Harvey, A.P.; Harvey, K.Y.; Nosalski, R.; Anyfanti, P.; Camargo, L.L.; Lacchini, S.; Ryazanov, A.G.; Ryazanova, L.; et al. Chanzyme TRPM7 protects against cardiovascular inflammation and fibrosis. Cardiovasc. Res. 2019. [Google Scholar] [CrossRef]
  220. Ryazanova, L.V.; Rondon, L.J.; Zierler, S.; Hu, Z.; Galli, J.; Yamaguchi, T.P.; Mazur, A.; Fleig, A.; Ryazanov, A.G. TRPM7 is essential for Mg(2+) homeostasis in mammals. Nat. Commun. 2010, 1, 109. [Google Scholar] [CrossRef]
  221. Du, J.; Silverman, D.; Liang, B.; Yue, L. Ca2+-permeable TRP channels in human cardiac fibroblasts. Circulation 2007, 116 (Suppl. 16), 188. [Google Scholar]
  222. Qin, X.; Yue, Z.; Sun, B.; Yang, W.; Xie, J.; Ni, E.; Feng, Y.; Mahmood, R.; Zhang, Y.; Yue, L. Sphingosine and FTY720 are potent inhibitors of the transient receptor potential melastatin 7 (TRPM7) channels. Br. J. Pharmacol. 2013, 168, 1294–1312. [Google Scholar] [CrossRef] [Green Version]
  223. Zhang, Y.J.; Ma, N.; Su, F.; Liu, H.; Mei, J. Increased TRPM6 expression in atrial fibrillation patients contribute to atrial fibrosis. Exp. Mol. Pathol. 2015, 98, 486–490. [Google Scholar] [CrossRef]
  224. Jiang, J.; Li, M.; Yue, L. Potentiation of TRPM7 inward currents by protons. J. Gen. Physiol. 2005, 126, 137–150. [Google Scholar] [CrossRef] [PubMed]
  225. Bers, D.M. Calcium fluxes involved in control of cardiac myocyte contraction. Circ. Res. 2000, 87, 275–281. [Google Scholar] [CrossRef]
  226. Stafford, N.; Wilson, C.; Oceandy, D.; Neyses, L.; Cartwright, E.J. The Plasma Membrane Calcium ATPases and Their Role as Major New Players in Human Disease. Physiol. Rev. 2017, 97, 1089–1125. [Google Scholar] [CrossRef] [PubMed]
  227. Mohamed, T.M.A.; Abou-Leisa, R.; Stafford, N.; Maqsood, A.; Zi, M.; Prehar, S.; Baudoin-Stanley, F.; Wang, X.; Neyses, L.; Cartwright, E.J.; et al. The plasma membrane calcium ATPase 4 signalling in cardiac fibroblasts mediates cardiomyocyte hypertrophy. Nat. Commun. 2016, 7, 11074. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram illustrating cardiac fibrogenesis cascade and fibrosis-associated heart diseases. Pathological stresses stimulate fibroblasts to differentiate into myofibroblasts, during which an increase in intracellular Ca2+ plays a key role. Myofibroblasts synthesize and secrete extracellular matrix (ECM) proteins, matrix metalloproteinases (MMPs), and cytokines such as TGFβ1. Excessive deposition of ECM proteins results in cardiac fibrosis. This fibrogenesis cascade is perpetuated by TGFβ1 produced by myofibroblasts. Cardiac fibrosis is involved in a variety of pathological remodeling, which can lead to arrhythmia, hypertrophy and heart failure.
Figure 1. Schematic diagram illustrating cardiac fibrogenesis cascade and fibrosis-associated heart diseases. Pathological stresses stimulate fibroblasts to differentiate into myofibroblasts, during which an increase in intracellular Ca2+ plays a key role. Myofibroblasts synthesize and secrete extracellular matrix (ECM) proteins, matrix metalloproteinases (MMPs), and cytokines such as TGFβ1. Excessive deposition of ECM proteins results in cardiac fibrosis. This fibrogenesis cascade is perpetuated by TGFβ1 produced by myofibroblasts. Cardiac fibrosis is involved in a variety of pathological remodeling, which can lead to arrhythmia, hypertrophy and heart failure.
Jcdd 06 00034 g001
Figure 2. Ca2+ signaling mechanisms in cardiac fibroblasts and myofibroblasts. Intracellular Ca2+ levels are finely controlled by: (1) Ca2+ entry through Ca2+-permeable channels in the plasma membrane, including TRP channels, P2X receptors, and Orai/STIM channels; (2) Ca2+ release via IP3Rs in the ER; and (3) Ca2+ extrusion pumps, including SERCA in the ER and PMCA4 in the plasma membrane. Pathophysiological stimuli can activate Gq-coupled receptors to induce Ca2+ release, which is secondarily followed by Ca2+ entry. Receptor stimulation can also directly activate Ca2+-permeable (TRP) channels in the plasma membrane. Other ion channels, including voltage-gated K+ (Kv) channels, Kir channels, Ca2+-activated potassium (KCa) channels, and NaV channels, may influence the resting membrane potential or depolarization to indirectly influence Ca2+ entry in fibroblasts and myofibroblasts. An increase in intracellular Ca2+ activates the calcineurin/NFAT (CN/NFAT), ERK1/2, ROS/RhoA, and sFRP2 pathways to promote profibrotic gene expression.
Figure 2. Ca2+ signaling mechanisms in cardiac fibroblasts and myofibroblasts. Intracellular Ca2+ levels are finely controlled by: (1) Ca2+ entry through Ca2+-permeable channels in the plasma membrane, including TRP channels, P2X receptors, and Orai/STIM channels; (2) Ca2+ release via IP3Rs in the ER; and (3) Ca2+ extrusion pumps, including SERCA in the ER and PMCA4 in the plasma membrane. Pathophysiological stimuli can activate Gq-coupled receptors to induce Ca2+ release, which is secondarily followed by Ca2+ entry. Receptor stimulation can also directly activate Ca2+-permeable (TRP) channels in the plasma membrane. Other ion channels, including voltage-gated K+ (Kv) channels, Kir channels, Ca2+-activated potassium (KCa) channels, and NaV channels, may influence the resting membrane potential or depolarization to indirectly influence Ca2+ entry in fibroblasts and myofibroblasts. An increase in intracellular Ca2+ activates the calcineurin/NFAT (CN/NFAT), ERK1/2, ROS/RhoA, and sFRP2 pathways to promote profibrotic gene expression.
Jcdd 06 00034 g002
Table 1. Membrane currents in cardiac fibroblasts.
Table 1. Membrane currents in cardiac fibroblasts.
Types of Ion Channels Names of the Ion ChannelsCell Types Expressing the Ion Channels
Voltage-gated Na+ currentsINa.TTX, INa.TTXR, [51]; INa.TTXR [61] NaV1.2, NaV1.9 [62]Human cardiac fibroblasts [51,63]
Voltage-gated K+ channelsTransient outward K+ current, Ito [51] [47,64]Human cardiac fibroblasts [51], neonatal rat cardiac fibroblasts [47,64]
Delayed rectifier K+ currents: IK [48,52], IKf and IKS [47], IKDR [51] Neonatal [47,64] and adult rat cardiac fibroblasts [48,52]; human cardiac fibroblasts [51]
Inward-rectifier K+ channelsInward-rectifying K+ currents: Kir [51,52], IK1 [44,53]Human cardiac fibroblasts [51], adult rat cardiac fibroblasts [52], dog atrial fibroblasts [44,53]
ATP-activated K+ channelsKATP [54,56]Rat ventricular fibroblasts [54,56]
Ca2+-activated K+ channelsBig conductance K+ currents activated by Ca2+: BKCa [51,57]Human cardiac fibroblasts [51,57]
Intermediate conductance K+ currents activated by Ca2+: KCa3.1 [59], IKCa [65]Rat adult cardiac fibroblasts [59,65]
Cl channelsSwelling-induced Cl currents, ICl-swell [51]Human cardiac fibroblasts [51]
Voltage-gated H+ channelsVoltage-gated H+-currents [66]Human cardiac fibroblasts [66]
Store-operated Ca2+ channels (Orai/STIM)Ca2+ release-activated Ca2+ currents: ICRAC [67,68]Human ventricular fibroblasts [67,68]
TRP channelsTRPC-like currents (TRPC3 or TRPC6) [69]Adult rat ventricular fibroblasts [69]
TRPC3 [70]Rat atrial fibroblasts [70]
TRPV4 [71]Rat cardiac fibroblasts [71]
TRPM2 [59]Rat adult fibroblasts [59]
TRPM7 [37,72]Mouse, rat, and human cardiac fibroblasts [37,72]
TRPA1 [73]Human cardiac fibroblasts [73]
Table 2. Potential role of Ca2+-permeable TRP channels in cardiac fibroblasts and fibrosis-associated heart diseases.
Table 2. Potential role of Ca2+-permeable TRP channels in cardiac fibroblasts and fibrosis-associated heart diseases.
Channels Expression in Cardiac CellsCellular Function in FibroblastsMethods Used for In Vivo Pathological Functions Potential Roles in Heart Diseases References
TRPC3 Myocytes, fibroblasts Proliferation, differentiationSpecific blockersArrhythmia, hypertrophy, heart failure[70,166,167,168,169,170]
TRPC6Myocytes, fibroblastsProliferation, differentiation Specific blockers; gene knockoutArrhythmia, hypertrophy, heart failure[38,164,171,172]
TRPV1Fibroblasts Specific activators and inhibitorsHypertrophy, heart failure[173,174,175,176,177]
TRPV2Myocytes, fibroblastsDifferentiationInhibitors; gene knockoutDilated and ischemic cardiomyopathy [178,179,180,181]
TRPV3Myocytes, fibroblasts ProliferationActivatorsHypertrophy[182,183]
TRPV4Myocytes, fibroblasts DifferentiationInhibitors; gene knockoutMyocardial infarction[39,184,185]
TRPM2Myocytes, fibroblasts Gene knockoutIschemic injury[186,187,188,189]
TRPM7Myocytes, fibroblasts Proliferation, differentiationshRNACardiac fibrosis[37,72,190,191,192,193,194,195,196]
TRPA1Myocytes, fibroblastsDifferentiationInhibitors, gene knockoutHypertrophy, heart failure[197,198,199,200]

Share and Cite

MDPI and ACS Style

Feng, J.; Armillei, M.K.; Yu, A.S.; Liang, B.T.; Runnels, L.W.; Yue, L. Ca2+ Signaling in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases. J. Cardiovasc. Dev. Dis. 2019, 6, 34. https://doi.org/10.3390/jcdd6040034

AMA Style

Feng J, Armillei MK, Yu AS, Liang BT, Runnels LW, Yue L. Ca2+ Signaling in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases. Journal of Cardiovascular Development and Disease. 2019; 6(4):34. https://doi.org/10.3390/jcdd6040034

Chicago/Turabian Style

Feng, Jianlin, Maria K. Armillei, Albert S. Yu, Bruce T. Liang, Loren W. Runnels, and Lixia Yue. 2019. "Ca2+ Signaling in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases" Journal of Cardiovascular Development and Disease 6, no. 4: 34. https://doi.org/10.3390/jcdd6040034

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