Next Article in Journal
A Review of CAR-T Combination Therapies for Treatment of Gynecological Cancers
Previous Article in Journal
Structural Evaluation of a Nitroreductase Engineered for Improved Activation of the 5-Nitroimidazole PET Probe SN33623
Previous Article in Special Issue
Perspective and Therapeutic Potential of the Noncoding RNA–Connexin Axis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calcium Regulation of Connexin Hemichannels

by
Erva Bayraktar
1,2,†,
Diego Lopez-Pigozzi
1,2,† and
Mario Bortolozzi
1,2,3,*
1
Veneto Institute of Molecular Medicine (VIMM), Via Orus 2, 35129 Padova, Italy
2
Department of Physics and Astronomy “G. Galilei”, University of Padua, Via Marzolo 8, 35131 Padova, Italy
3
Institute of Endocrinology and Oncology “Gaetano Salvatore” (IEOS-CNR), Via Pietro Castellino 111, 80131 Napoli, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(12), 6594; https://doi.org/10.3390/ijms25126594
Submission received: 1 May 2024 / Revised: 7 June 2024 / Accepted: 12 June 2024 / Published: 15 June 2024

Abstract

:
Connexin hemichannels (HCs) expressed at the plasma membrane of mammalian cells are of paramount importance for intercellular communication. In physiological conditions, HCs can form gap junction (GJ) channels, providing a direct diffusive path between neighbouring cells. In addition, unpaired HCs provide conduits for the exchange of solutes between the cytoplasm and the extracellular milieu, including messenger molecules involved in paracrine signalling. The synergistic action of membrane potential and Ca2+ ions controls the gating of the large and relatively unselective pore of connexin HCs. The four orders of magnitude difference in gating sensitivity to the extracellular ([Ca2+]e) and the cytosolic ([Ca2+]c) Ca2+ concentrations suggests that at least two different Ca2+ sensors may exist. While [Ca2+]e acts as a spatial modulator of the HC opening, which is most likely dependent on the cell layer, compartment, and organ, [Ca2+]c triggers HC opening and the release of extracellular bursts of messenger molecules. Such molecules include ATP, cAMP, glutamate, NAD+, glutathione, D-serine, and prostaglandins. Lost or abnormal HC regulation by Ca2+ has been associated with several diseases, including deafness, keratitis ichthyosis, palmoplantar keratoderma, Charcot–Marie–Tooth neuropathy, oculodentodigital dysplasia, and congenital cataracts. The fact that both an increased and a decreased Ca2+ sensitivity has been linked to pathological conditions suggests that Ca2+ in healthy cells finely tunes the normal HC function. Overall, further investigation is needed to clarify the structural and chemical modifications of connexin HCs during [Ca2+]e and [Ca2+]c variations. A molecular model that accounts for changes in both Ca2+ and the transmembrane voltage will undoubtedly enhance our interpretation of the experimental results and pave the way for developing therapeutic compounds targeting specific HC dysfunctions.

1. Introduction

Connexins are a family of tetraspan membrane proteins encoded by 21 genes in humans [1]. The oligomerization of different connexin isoforms into hexameric structures, called connexons or hemichannels (HCs), usually occurs during the transition from the rough endoplasmic reticulum to the Golgi apparatus [2]. Only a fraction of the total connexin expressed by the cell reaches the plasma membrane [3,4], where two HCs belonging to adjacent membranes can dock end–to–end to form a junctional channel. Tens to thousands of junctional channels can line up in a dense hexagonal pattern called a gap junction (GJ) [5]. Various compounds with a molecular mass of up to approximately 1 kDa, including water molecules, ions, second messengers, amino acids, nucleotides, and glucose, can be exchanged by passive diffusion through GJ channels that connect the cytoplasms of neighbouring cells or even different regions of the same cell, as in the case of the myelin sheath [6]. While connexin HCs are normally closed at rest, they can open under physiological conditions, thus allowing sustained ion fluxes and permeation of messenger molecules [7].
Adenosine triphosphate (ATP) is a fundamental signalling molecule that can be released from the cytosol by connexin HCs, including those formed by connexin 26 (Cx26) [8,9], Cx30 [10,11], Cx30.2/Cx31.3 [12], Cx32 [13], Cx36 [14], Cx38 [15], Cx40 [16], and Cx43 [17,18]. The ATP released through connexin HCs may promote cytosolic Ca2+ oscillations and intercellular Ca2+ wave propagation by the activation of purinergic receptors [16,17,19,20,21,22] (Figure 1). Other key messenger molecules released by connexin HCs include cyclic adenosine monophosphate (cAMP) [23], glutamate [24,25], oxidized nicotinamide adenine dinucleotide (NAD+) [26,27], glutathione [28], D-serine [29], and prostaglandins [30,31]. These molecules can mediate autocrine and paracrine processes that regulate the development, homeostasis, function, and regeneration of several organs.
Connexin HCs have been found to be involved in cell proliferation [32], cell–cell adhesion [33], light processing by the retina [34], metabolic homeostasis, and transparency of the lens [35,36]. They are also involved in hearing maturation and function [37,38], bone growth, remodelling and protection against oxidative stress [39,40,41,42], isosmotic regulation of cell volume [43], salt and water reabsorption in the renal tubule [11], blood–brain barrier permeability [44], and the development and functionality of the peripheral and central nervous system [29,45,46,47,48,49,50]. Connexin HCs also play a role in autophagy [51], tumour growth [52], mitochondrial permeability in cardiac cells [53,54], atrioventricular conduction [55,56], vessel contractility [57], and epidermal barrier homeostasis [58].
The opening and closure of connexin HCs are regulated by several physiological parameters, such as transmembrane voltage [59,60], cell redox potential [61], phosphorylation [62,63], membrane mechanical stretch [64,65], amino sulfonates [66], CO2 [67], pH [68,69], and cations [70,71], including extracellular [72,73] and intracellular Ca2+ [13,74,75].
The complex interaction of Ca2+ with connexin HCs is strictly interconnected with many biological functions accomplished by this ion in the cell. This review focuses on HC regulation by Ca2+, one of the most important factors controlling molecular fluxes through connexin HCs in health and disease.

2. Extracellular Ca2+ Regulation of Connexin HCs

In humans, the extracellular concentration of Ca2+ ([Ca2+]e) is found in a narrow range of 1.1–1.4 mM in most organs [76], with the notable exception of the cochlear endolymph [77]. A reduction in [Ca2+]e to hundreds or tens of micromolars typically causes a marked increase in the amplitude of HC currents, thus shifting their activation to more negative potentials and altering the activation and deactivation kinetics [73]. Low values of [Ca2+]e increase the activity of most, if not all, connexin HCs, including Cx26 [78], Cx30 [60], Cx30.2/Cx31.3 [12], Cx32 [72], Cx37 [79], Cx39 [80], Cx40 [81], Cx43 [63], Cx45 [82], Cx46 [83], and Cx50 [84] (Table 1). It is important to note that connexin HCs cannot be considered as regulators of [Ca2+]e, since their opening does not activate any feedback mechanism for [Ca2+]e adjustment. Instead, [Ca2+]e variations modulate the HC opening probability, which is key for several downstream physiological mechanisms.
In the intact central nervous system, [Ca2+]e can be reduced from 1.2–1.4 mM to less than 0.7 mM during periods of intense neuronal activity, a phenomenon that activates purinergic feedback signalling from astrocytes to interneurons mediated by Cx43 HCs [86]. In non-sensory cells of the cochlea, the low endolymphatic [Ca2+]e (20–30 μM [77]) biases Cx26/Cx30 HCs towards the open state, thus altering the ATP-dependent intercellular Ca2+ signalling that is required for refinement of afferent innervations of outer hair cells [87]. In the mammalian epidermis, a characteristic [Ca2+]e gradient between lower and upper layers plays a crucial role in the processes of keratinocyte differentiation and formation of the epidermal permeability barrier [88].
Specific information about the dependency of the HC gating on [Ca2+]e is available for several connexin isoforms, as detailed in the following.
Cx26: Atomic force microscopy (AFM) experiments performed in isolated membranes of HeLa cells expressing rat Cx26 (rCx26) showed a reduction in the HC extracellular inner diameter from 1.3 to 0.5 nm by adding 0.5 mM Ca2+ to the [Ca2+]e-free solution [89]. In Xenopus oocytes expressing human Cx26 (hCx26), the HC current dependent on [Ca2+]e ranged from a maximum at 0.01 mM [Ca2+]e to a minimum value at 10 mM [Ca2+]e, with an EC50 around 0.25 mM [78]. Additionally, lowering [Ca2+]e from 0.75 mM to 0.1 mM increased the HC current by approximately 85% at +40 mV, while increasing [Ca2+]e from 0.75 mM to 3.5 mM induced a decrease of about 80% [90]. A close examination of the HC structure revealed that the negatively charged residue at position 50 (D50) of hCx26 is critical for HC gating and, together with the E47, might directly interact with Ca2+ ions to induce occlusion of the pore [73,78,91]. The D50 and E47 residues were proposed to play a role in stabilizing the HC open state in low-Ca2+ conditions, forming salt bridges with the K61 and R75-R184 residues, respectively. Bridge disruption at high [Ca2+]e would destabilize the open state, thus facilitating HC closure. The conserved glycine at position 45 of hCx26, hCx30, hCx32, and hCx43 HCs was also found to be an integral part of the [Ca2+]e sensor [92].
Cx32: In Xenopus oocytes expressing hCx32 HCs, HC currents ranged from a maximum at 0.5 mM [Ca2+]e to a minimum value at 5 mM [Ca2+]e, with an EC50 around 1.3 mM [72]. This very high EC50 value might be due to the application of a strong (+80 mV) voltage stimulation that favoured the opening of Cx32 HCs during the current recording. The effect of voltage on the HC current dependence on [Ca2+]e is well documented for another connexin (Cx46), whose EC50 shifted from 0.08 mM [Ca2+]e at values ≤ −20 mV to 0.5 mM [Ca2+]e at +20 mV [93]. Based on molecular dynamics simulations [94], this EC50 shift might be simply explained by a change in the HC binding affinity to Ca2+. Two Asp residues, D169 and D178, were found to be implicated in the Ca2+-induced blockage and conductance properties of hCx32 HCs [72]. The characteristic transitions attributable to the fast gate between a main open state of high conductance (90 pS) and a residual open state (18 pS) only occurred at low [Ca2+]e and upon hyperpolarization. Gomez-Hernandez et al. [72] also found that substitution of Ca2+ in the external solution with other divalent cations inhibited hCx32 HC activation. The potential of different divalent cations to induce such a block followed the sequence: Cd2+ > Co2+ ≈ Ca2+ > Mg2+ > Ba2+.
Cx37: In Xenopus oocytes expressing hCx37 HCs, the HC current dependent on [Ca2+]e ranged from a maximum at 0.02 mM [Ca2+]e to a minimum value at 1 mM [Ca2+]e, with an EC50 of about 0.1 mM [79]. In rat brain endothelial cells (RBE4) expressing endogenous Cx37, as well as in HeLa cells expressing exogenous hCx37, dye uptake and release through Cx37 HCs were strongly inhibited by the GAP27 peptide, and single-channel electrophysiological studies indicated that GAP27 inhibits unitary HC currents [44].
Cx43: In rat Novikoff hepatoma cells expressing endogenous Cx43, a reduction in extracellular Ca2+ but not Mg2+ was a key factor for HC opening and dye uptake. An increased uptake started at 1 mM [Ca2+]e, reaching a maximal level at 10 μM [Ca2+]e [63]. In Xenopus oocytes expressing hCx43, lowering the external divalent cation concentration (Ca2+ and Mg2+ free) decreased the resting potential and the input resistance. Both parameters recovered their initial values upon restoring the [Ca2+]e to normal millimolar values [95]. In HeLa cells expressing rCx43, HC opening increased only modestly at positive potentials in zero Ca2+ or zero Ca2+-EGTA solutions [59].
Cx45: In HeLa cells expressing mouse Cx45 (mCx45), the HC current dependency on [Ca2+]e was a combination of two Hill equations with different sensitivities (K1 = 0.66 μM and K2 = 216 μM) and a very low EC50 (~1 μM) [82]. It may be argued that the dual sensitivity reflects the coexistence of both the exogenous mCx45 and the endogenous hCx45 [96].
Cx46: In Xenopus oocytes, rCx46 HCs have a slightly lower apparent affinity for Ca2+ than hCx26 HCs (KD 0.6 mM vs. 0.33 mM) [73]. Regarding Ca2+ sensitivity, the E48 and D51 residues of Cx46 played a role like the analogous residues E47 and D50 in Cx26. In the same cellular system [83], L35 also appeared as an important residue for HC closure by Ca2+, and single HC recordings suggested that divalent cations act as stabilizers of the fully closed conformation rather than as gating particles [97]. In Xenopus oocytes expressing hCx46, lowering [Ca2+]e from 0.1 mM to 0.01 mM increased the HC current by approximately 150% at +20 mV, whereas increasing [Ca2+]e from 0.1 mM to 1 mM induced a decrease of about 80% [71]. Like other connexins, a high [Mg2+]e also acted as a blocker but to a much lower extent. The experiments by Ebihara et al. [71] applying sequential combinations of different [Ca2+]e and [Mg2+]e values suggested that there are at least two distinct binding sites for divalent cations that have different relative affinities for Ca2+ and Mg2+ and modulate different steps in the gating process.
Cx50: In HeLa cells pre-loaded with Lucifer Yellow (LY) dye and expressing hCx50, removal of [Ca2+]e (2 mM) increased the rate of dye leakage and activated a voltage-dependent outward transmembrane current [84]. Notably, Cx50 and Cx46 share very similar amino acid sequences and HC sensitivity to [Ca2+]e. In Xenopus oocytes expressing mCx50 or rCx46, [Ca2+]e shifted the I-V curve of Cx46 but not that of Cx50 HCs [98]. Cx50 HCs are also much more sensitive to external pH than Cx46 HCs. Interestingly, the replacement of extracellular Na+ with K+ or other monovalent cations while maintaining a constant high [Ca2+]e resulted in 10-fold potentiation of mCx50 HC currents, which reversed upon restoring a normal Na+ concentration [70]. In contrast, rCx46 HCs exhibited a modest increase upon substituting Na+ with K+. The primary effect of K+ appeared to be a reduction in the ability of Ca2+, as well as other divalent cations, to close Cx50 HCs due to the specific connexin sequence.

3. Cytosolic Ca2+ Regulation of Connexin HCs

The [Ca2+]c of most eukaryotic cells is maintained at ~100 nM in resting conditions, a value 10,000-fold lower than the extracellular concentration [99]. Rapid variations in [Ca2+]c in the nanomolar or micromolar range, triggered by different extracellular and intracellular stimuli, are sufficient to activate and control several cellular mechanisms alone, such as secretion, gene expression, muscle contraction, metabolism, and also connexin channel gating [72,100,101]. The increase in [Ca2+]c typically results from either the influx of extracellular Ca2+ via plasma membrane Ca2+ channels or the release of Ca2+ from internal stores via InsP3Rs and ryanodine receptors (RyRs). Connexin HCs that show an increased open probability upon [Ca2+]c variation include Cx26 [102], Cx30 [60], Cx32 [13], Cx43 [75], Cx45 [103], Cx46 [85] (Table 1), and most likely also Cx37 [44] and Cx40 [16]. In the organ of Corti, ATP release through Cx26 and Cx30 HCs is triggered by spontaneous [Ca2+]c oscillations in supporting cells and contributes to the propagation of the intercellular Ca2+ signalling that is responsible for synaptic refinement [8,87]. In astrocytes, a migratory phenotype is acquired under pro-inflammatory conditions by a [Ca2+]c-dependent opening of Cx43 HCs, followed by release of ATP, activation of the P2X7 receptor, and Ca2+ influx [104]. Retinal pigment epithelium (RPE) cells were found to release ATP through Cx43 HCs that increased proliferation and stimulated DNA synthesis in neural retinal progenitor cells [105]. Upon mechanical stress, an increased [Ca2+]c elicited by the activation of Piezo1 channels opens Cx43 HCs through PI3K-Akt, which in turn leads to bone anabolic function [42]. Furthermore, NAD+, prostaglandin E2, and ATP release by Cx43 HCs in response to oxidative stress was shown to serve as a protective mechanism for osteocytes, which may accumulate reactive oxygen species with skeletal aging [39].
Specific information about the dependency of the HC gating on [Ca2+]c is available for several connexin isoforms, as detailed in the following.
Cx26: In HeLa cells expressing rCx26, a [Ca2+]c increase mediated by linoleic acid (LA) appeared to be essential to enhance Cx26 HC activity [102]. The synthetic Ca2+ chelator BAPTA and PI3K/Akt inhibitors were found to reduce ethidium (Etd) bromide uptake through Cx26 HCs, suggesting that the LA effect was mediated by the increase in [Ca2+]c and activation of the PI3K/Akt-dependent pathway. Notably, an LA-mediated [Ca2+]c increase stimulated Etd uptake also through mCx32, mCx43, and mCx45 HCs, while drastically reducing rCx26 GJ-mediated dye coupling [102]. In mouse cochlear organotypic cultures, focal InsP3 photoliberation elicited a [Ca2+]c response in the irradiated supporting cells that peaked at around 500 nM and spread radially to several orders of unstimulated cells [8]. mCx26 and mCx30 HCs contributed to the propagation of this intercellular Ca2+ wave by releasing ATP, without the contribution of P2X7R or pannexin 1. ATP release was not observed in Cx26 KO or Cx30 KO cultures upon a [Ca2+]c increase triggered by InsP3 photoliberation.
Cx32: In bladder cancer epithelial cells (ECV304) and C6 glioma cells expressing hCx32, Leybaert and co-workers [13] showed that (i) an InsP3-mediated increase in [Ca2+]c is sufficient to trigger the opening of Cx32 HCs, promoting dye uptake and ATP release, and (ii) the HC open probability has a bell-shaped dependency peaking at 500 nM [Ca2+]c. Both small and large Ca2+ stimuli were ineffective in opening Cx32 HCs, and only [Ca2+]c changes between 200 nM and 1000 nM were successful. Cytosolic photorelease of Ca2+ from o-nitrophenyl EGTA (NP-EGTA) or activation of Ca2+ influx by the A23187 ionophore also triggered ATP release in a dose-dependent manner, with ATP responses disappearing at stronger stimulation [13]. In single HeLa cells expressing hCx32, a local puff of extracellular ATP or histamine stimulated a [Ca2+]c increase that triggered the opening of Cx32 HCs held at a −20 mV transmembrane potential [74]. The temporal dynamics of HC opening and closure were quantified in terms of membrane conductance variations with an exponential behaviour with rise (opening) and decay (closure) times of around 1 s.
Cx43: In C6 glioma and HeLa cells stably transfected with Cx43, the application of different concentrations of the Ca2+ ionophore 4-Br-A23187 (calcimycin) induced [Ca2+]c transients and ATP release with a bell-shaped dependence on [Ca2+]c which was maximal at ~500 nM [75]. Inhibition of ATP release was achieved in a concentration-dependent way by both GAP26 and GAP27, two peptides that mimic a short sequence on the first and second extracellular loops of Cx43, respectively. The peptide GAP19, which mimics a short sequence of the Cx43 cytoplasmic loop (CL), also inhibited [Ca2+]c-triggered ATP release in a concentration-dependent manner without affecting the GJ conductance and dye exchange [106]. In HeLa cells, an elevation in [Ca2+]c from 50 to 200–500 nM in the absence of an electrical trigger was ineffective in opening single Cx43 HCs, but lowered the voltage threshold for HC opening by ~15 mV and potentiated the unitary HC current [107]. The HCs closed with a further elevation in [Ca2+]c to 1 μM, exhibiting a bell-shaped dependence of the open probability on [Ca2+]c peaking at 500 nM and that was very similar to the previous indirect measurement for Cx43 expressed in intact cells [75]. In isolated pig ventricular cardiomyocytes, [Ca2+]c elevation to 500 nM potentiated the HC current, which was inhibited by GAP26 and GAP27. A similar study was performed in astrocytes of the prefrontal cortex of newborn mice, where mCx43 HC activity was found to be dependent on [Ca2+]c and associated with D-serine release [50]. In rat brain endothelial cell lines expressing rCx43, single-cell photoactivation of InsP3 triggered an intercellular purinergic signalling which is [Ca2+]c-dependent and can be blocked by GAP26 [108]. The opening of Cx43 HCs was evoked by spontaneous elevations in [Ca2+]c during Ca2+ waves in trigger cells of the chick RPE and was blocked by GAP26 [105]. The concept of trigger cells was demonstrated by luminometric extracellular ATP imaging, which showed that ATP is released as a burst from a point source of the cell [109]. Interestingly, Cx43 HC opening induced by oxidative stress in osteocytes was inhibited by the depletion in [Ca2+]c with BAPTA-AM, suggesting that [Ca2+]c triggers the activity of mCx43 HCs [41]. Conversely, blockade of HCs with a Cx43 antibody did not affect [Ca2+]c [39].
Cx45: In HeLa cells expressing mCx45, HC openings were triggered at 100 nM [Ca2+]c in the patch pipette [82]. In the same expression system, a progressive [Ca2+]c rise stimulated by the application of 2.5 μM calcimycin increased Etd uptake within the cell [103]. The non-specific HC blocker La3+ significantly reduced this uptake.
Cx46: In HeLa cells stably transfected with hCx46, a [Ca2+]c transient mediated by 2.5 μM of the Ca2+ ionophore ionomycin led to a dramatic increase in Etd uptake by Cx46 HCs [85]. Replacement of extracellular Na+ with K+ led to cell depolarization but reduced Etd uptake.

4. Pathological Alterations of HC Gating by Ca2+

Lost or abnormal HC activity at the plasma membrane has been associated with inflammatory conditions [110,111] and inherited diseases, like muscular dystrophy [112], syndromic and non-syndromic deafness [113], keratitis and hystrix-like ichthyosis deafness (KID/HID) syndrome [114], the X-linked form of Charcot–Marie–Tooth neuropathy (CMT1X) [115], oculodentodigital dysplasia (ODDD) [116], keratoderma–hypotrichosis–leukonychia totalis syndrome (KHLS) [117], erythrokeratodermia variabilis (EKV) [118], and congenital cataracts [119]. A single amino acid substitution in the connexin sequence can severely affect the correct HC function, leading to uncontrolled ionic leakage and release of molecules altering extracellular signalling pathways and potentially toxic for neighboring cells [120,121,122,123,124]. The biophysical properties altered in mutant HCs expressed at the plasma membrane may relate to their density, the pore selective permeability, or the gating mechanisms, including Ca2+-dependent regulation [73]. Several pathological mutations have been linked to altered sensitivity to extracellular and cytosolic Ca2+ ions, resulting in changes in the normal HC activity (Table 2). As a general criterion, the sensitivity to Ca2+ can be considered altered when an HC defect is present in normal [Ca2+]e, or there is a rightward or leftward shift in the HC current dependence on [Ca2+]e. Some mutations lack a clear link between HC dysfunction and Ca2+ deregulation. The mere observation of an increased HC activity in zero [Ca2+]e solution, which shifts the HC to the fully open state, can also be attributed to alteration of other HC properties, such as pore permeability or voltage-dependent gating.
Mutations of the GJB2 gene, which encodes Cx26, were linked to sensorineural hearing loss, keratitis, and severe skin lesions [125,126]. In particular, G11E, G12R, N14K, N14Y, I30N, A40V, G45E, E47Q, D50N/Y/A, and A88V mutations were associated with a reduced HC closure by [Ca2+]e or an increased permeability to Ca2+ ions, which resulted in leaky HCs and compromised keratinocyte viability in vitro [78,123,127,128,129,130,131,132,133,134]. Instead, S17F, N54K, R75W, and S183F pathological mutations were associated with a decreased HC activity [123,129,130,135]. Examination of the skin of KID transgenic mice showed alterations in the epidermal [Ca2+]e gradient that were correlated with altered lipid secretion and defects in the epidermal water barrier [133,136]. Negatively charged residues (D46, E47, Q48, D50, K61, R184) lining the Cx26 HC pore are critical for the gating dependency on [Ca2+]e, but they do not form the HC gate [73,137].
Table 2. Impact of connexin mutations on the HC functionality under different [Ca2+]e and [Ca2+]c conditions. Normal [Ca2+]e solution refers to an extracellular solution with a Ca2+ concentration in the mM range (typically 2–5 mM), whereas a [Ca2+]e-free solution is prepared without Ca2+ or with Ca2+ buffered by EGTA. Abbreviations: connexin (Cx), N-terminus (NT), transmembrane domain (TM), extracellular loop (EL), cytoplasmic loop (CL), C-terminus (CT); erythrokeratodermia variabilis (EKV), keratitis ichthyosis deafness (KID) syndrome, keratoderma–hypotrichosis–leukonychia totalis syndrome (KHLS), oculodentodigital dysplasia (ODDD).
Table 2. Impact of connexin mutations on the HC functionality under different [Ca2+]e and [Ca2+]c conditions. Normal [Ca2+]e solution refers to an extracellular solution with a Ca2+ concentration in the mM range (typically 2–5 mM), whereas a [Ca2+]e-free solution is prepared without Ca2+ or with Ca2+ buffered by EGTA. Abbreviations: connexin (Cx), N-terminus (NT), transmembrane domain (TM), extracellular loop (EL), cytoplasmic loop (CL), C-terminus (CT); erythrokeratodermia variabilis (EKV), keratitis ichthyosis deafness (KID) syndrome, keratoderma–hypotrichosis–leukonychia totalis syndrome (KHLS), oculodentodigital dysplasia (ODDD).
Cx IsoformMutationCx DomainMutant HC Defective PropertiesLinked Disease
Cx26G11ENTIncreased Ca2+ leakage in normal [Ca2+]e [127].KID syndrome
G12RNTIncreased dye uptake in both normal and [Ca2+]e-free solutions, halted [Ca2+]e-dependent deactivation kinetics, increased Ca2+ leakage [123,128,129].Syndromic deafness
G12VNTIncreased dye uptake in [Ca2+]e-free solution [129].Non-syndromic deafness
N14KNTReduced [Ca2+]e sensitivity leading to increased HC currents and slowed deactivation kinetics, increased Ca2+ leakage [123,130,131].Clouston syndrome/KID syndrome
N14YNTIncreased dye uptake and HC currents in both normal and [Ca2+]e-free solutions, increased Ca2+ leakage [129].Syndromic deafness
S17FNTDecreased dye uptake in both normal and [Ca2+]e-free solutions, leaky HCs when co-expressed with Cx30 [123,129,138].KID syndrome
I30NTM1Increased dye uptake in both normal and [Ca2+]e-free solutions, increased Ca2+ leakage [132].KID syndrome
V37ITM1Abolished dye uptake in [Ca2+]e-free solution [139].Deafness
A40GTM1Abolished dye uptake in [Ca2+]e-free solution [139].Deafness
A40VTM1Reduced [Ca2+]e sensitivity leading to increased HC currents [133].KID syndrome
G45EEL1Increased dye uptake at a normal [Ca2+]e and a reduced [Ca2+]e sensitivity, leading to increased HC currents and Ca2+ leakage [133,140].KID syndrome
D46CEL1Halted [Ca2+]e-dependent deactivation kinetics [73].No association
E47KEL1Halted dye uptake in [Ca2+]e-free solution [140].Deafness
E47QEL1Reduced [Ca2+]e sensitivity, leading to increased HC currents and halted deactivation kinetics [73].No association
Q48AEL1Increased [Ca2+]e sensitivity, leading to decreased HC currents and reduced deactivation kinetics [137].No association
D50AEL1Reduced [Ca2+]e sensitivity, leading to increased HC currents and halted deactivation kinetics [134,137].KID syndrome
D50NEL1Reduced [Ca2+]e sensitivity, leading to increased HC currents and halted deactivation kinetics [73,78,123,127,130].KID syndrome
D50YEL1Increased dye uptake in both normal and [Ca2+]e-free solutions, halted [Ca2+]e-dependent deactivation kinetics, increased Ca2+ leakage [78,132].KID syndrome
N54KEL1Decreased dye uptake in [Ca2+]e-free solution [130].Bart–Pumphrey syndrome
R75WTM2Decreased HC currents in both normal and [Ca2+]e-free solutions [135].Deafness
A88VTM2Increased HC currents in [Ca2+]e-free solution [134].KID syndrome
S183FEL2Halted dye uptake in [Ca2+]e-free solution [130].Palmoplantar keratoderma and hearing loss
R184KEL2Reduced [Ca2+]e sensitivity leading to both increased HC currents and increased deactivation kinetics [73].No association
Cx30G11RNTIncreased ATP release in normal [Ca2+]e solution [141].Clouston syndrome
G45EEL1Reduced [Ca2+]e sensitivity, leading to dye uptake in both normal and [Ca2+]e-free solutions [92].No association
A88VTM2Increased ATP release in both normal and [Ca2+]e-free solutions, increased leakage of Ca2+ and other ions [58,141].Clouston syndrome
Cx30.3G12DNTIncreased dye uptake in [Ca2+]e-free solution [118].EKV
T85PTM2Increased dye uptake in [Ca2+]e-free solution [118].EKV
F189YTM4Increased dye uptake in both normal and [Ca2+]e-free solutions [118].EKV
Cx32G45EEL1Increased dye uptake in [Ca2+]e-free solution [92].No association
S85CTM2Increased HC current in normal [Ca2+]e solution [142].CMT1X
D169NEL2Reduced [Ca2+]e sensitivity, leading to increased HC currents [72].No association
D178NEL2Reduced [Ca2+]e sensitivity, leading to increased HC currents [72].No association
D178YEL2Reduced [Ca2+]e sensitivity, leading to increased HC currents [72].CMT1X
R220XCTReduced [Ca2+]c sensitivity, leading to halted HC current [74].CMT1X
Cx43L7VNTIncreased ATP release in normal [Ca2+]e solution [116].ODDD
G8VNTIncreased HC current in [Ca2+]e-free solution [143].KHLS
Y17SNTDecreased dye uptake in [Ca2+]e-free solution [144].ODDD
G21RTM1Decreased dye uptake in [Ca2+]e-free solution [144].ODDD
I31MTM1Increased ATP release in [Ca2+]e-free solution [122].ODDD
A40VTM1Decreased dye uptake in [Ca2+]e-free solution [144].ODDD
A44VTM1Increased HC current in [Ca2+]e-free solution [143].EKV
G45EEL1Increased dye uptake in [Ca2+]e-free solution [92].No association
F52dupEL1Decreased dye uptake in [Ca2+]e-free solution [144].ODDD
L90VTM2Decreased dye uptake in [Ca2+]e-free solution [144].ODDD
I130TCLDecreased dye uptake in [Ca2+]e-free solution [144].ODDD
G138RCLIncreased ATP release in [Ca2+]e-free solution [122].ODDD
G143SCLIncreased ATP release in [Ca2+]e-free solution [122].ODDD
E227DCTIncreased HC current in [Ca2+]e-free solution [143].EKV
M239XCTAbolished ATP release in [Ca2+]e-free solution and upon increased [Ca2+]c [145].No association
Cx46D47NEL1Increased [Ca2+]e sensitivity, leading to decreased HC currents [73].No association
E48QEL1Reduced [Ca2+]e sensitivity, leading to increased HC currents [73].No association
D51NEL1Reduced [Ca2+]e sensitivity, leading to increased HC currents [73].No association
N63SEL1Abolished HC currents that were restored in [Ca2+]e-free solution [146].Congenital cataracts
G143RCLIncreased dye uptake in both normal and [Ca2+]e-free solutions, decreased dye uptake upon increased [Ca2+]c [85,147].Congenital cataracts
Since mutations in the GJB1 gene that encodes Cx32 were first reported in 1993 [148], more than 450 different mutations associated with the X-linked dominant form of CMT (CMT1X) neuropathy have been discovered [115]. In the PNS, Cx32 was found to selectively be expressed in non-compact myelin of Schwann cells, but its physiological role is still up for debate. The D178Y mutation in the second extracellular loop of Cx32 affects an Asp-178 residue that induces a complete [Ca2+]e deregulation of the HC activity [72]. Cx32 HCs carrying a pathological CT domain truncation (R220X) fail to open in response to a canonical InsP3-mediated signal transduction cascade that elevates the [Ca2+]c [74]. Interestingly, the gating function of Cx32-R220X HCs was restored by both the intracellular and extracellular application of the peptide GAP24 that mimics the Cx32 CL.
More than 70 dominant mutations, mostly autosomal, of the GJA1 gene that encodes Cx43, were linked to ODDD, a development disorder characterized by craniofacial and limb disorders [149]. Over one-third of the mutations are localized in the CL of Cx43. In stable cell lines expressing enhanced yellow fluorescent protein (eYFP)-tagged hCx43 [144], propidium iodide uptake experiments at a low [Ca2+]e, demonstrated that Y17S, G21R, A40V, F52dup, L90V, and I130T mutations are associated with a reduced HC function compared with the wild-type Cx43. Instead, HeLa cells expressing Cx43 carrying ODDD mutations I31M, G138R, and G143S displayed an altered HC function with increased ATP release under zero-[Ca2+]e conditions [122]. In patient-derived fibroblasts, the ATP defect of G138R and G143S HCs was not significant, whereas the L7V mutation was found to be leaky at normal [Ca2+]e values [116]. The downregulation of L7V HCs found in patients’ fibroblasts was proposed as a protective mechanism against cytotoxicity.
Mutations in the GJA3 and GJA8 genes coding for Cx46 and Cx50, respectively, are a major cause of congenital cataracts [150]. More than 40 cataract-associated mutations are located in transmembrane and extracellular loops of Cx46. Instead, the pathological mutation G143R is located in the CL of Cx46, increasing HC leakage in resting conditions and reducing HC opening upon [Ca2+]c stimulation by ionomycin [85,147]. E47 and D50 residues in Cx26 correspond to residues E48 and D51 in Cx46. In particular, mutations E47Q and D50N/Y/A in Cx26 and E48Q and D51N in Cx46 decrease the [Ca2+]e sensitivity (i.e., leaky HCs), while Q48A and D47N mutations increase the sensitivity in Cx26 and Cx46, respectively [73,78,137].

5. Discussion

Ca2+ ions are key regulators of connexin HC functionality in health and disease. The four orders of magnitude difference in gating sensitivity to Ca2+ between the extracellular and intracellular HC domains suggests the existence of at least two different Ca2+ sensors. Under cell resting conditions, the synergistic action of the two gates, in combination with the transmembrane voltage, keeps this relatively large and unselective pore fully closed, thus preventing cytotoxic ionic leakage. The fact that both an increased and a decreased HC Ca2+ sensitivity is linked to disease suggests that Ca2+ in healthy cells finely regulates normal HC function. While [Ca2+]e is a relatively stable physiological parameter dependent only on the cell layer, compartment, and organ, [Ca2+]c can change rapidly, thus triggering sudden HC opening and release of extracellular bursts of messenger molecules involved in paracrine signalling.
There is a general consensus on the gating by [Ca2+]e through the direct binding of Ca2+ ions to specific extracellular residues of the six connexins forming the HC. This Ca2+ ring was proposed to act as an electrostatic occlusion and a stabilizer of the closed HC conformation [93,94]. Accessibility studies and numerical simulations suggest that the electrostatic effect does not hinder the access of ions or small molecules to positions deeper into the pore [78,97,151], whereas this conclusion is not generally shared [94,152]. Fitting electrophysiological experiments with Cx46 HCs by a model allosterically coupling Ca2+ binding and voltage sensing indicates that Ca2+ ions act like stabilizers of the closed HC conformation [93]. Coarse-grained molecular dynamics simulations of the Cx26 HC confirmed that Ca2+ coordination within the extracellular vestibule inhibits the transition to a wider pore state that would favour permeation [153]. This transition was not found in the Cx31.3 HC structure, which did not significantly change in the presence or absence of Ca2+ ions [154]. In single Cx32 HCs, a high [Ca2+]e inhibited transitions from the closed to the fully open state in both depolarization and hyperpolarization conditions, thus allowing only transitions to a residual substate [72].
The model by Pinto et al. [93] considers only [Ca2+]e and voltage as key variables of HC opening, without including the gating by [Ca2+]c or other cytosolic chemical gating mechanisms. Indeed, in connexin isoforms such as Cx32 and Cx43, the voltage alone can trigger HC opening only at unphysiological values ≥ +40 mV [59,72,74,107]. Instead, a physiological [Ca2+]c rise alone triggered by InsP3 can stimulate HC opening [74,155] while dramatically increasing the voltage sensitivity [107].
The molecular mechanisms underlying the exquisite regulation of connexin HCs by nanomolar variations in [Ca2+]c are still under debate. The interaction of the CL with the CT domain and calmodulin (CaM) was found to control the gating of both connexin GJs [156,157,158,159] and HCs [13,74,145,160]. In the pathological G143R mutation of Cx46, HC dysfunction was attributed to an increased interaction of the CL with the CT domain and CaM, possibly caused by the loss of the CL α-helical structure [85]. Lack of the CT domain in Cx32 [74] and Cx43 [145] caused the HC to fail to open in response to a [Ca2+]c increase. The opening was restored upon application of peptides (GAP24 and TAT–Cx43CT10, respectively) that are able to interact with the CL domain. Leybaert and coworkers [161] proposed that loss of the CL-CT domain interaction by a Ca2+-dependent activation of the actomyosin contractile system underlies the Cx43 HC closure mechanism upon [Ca2+]c overload (above 500 nM). Peracchia et al. [159] have proposed that CaM acts both as a Ca2+ sensor and a cork at the cytoplasmic mouth of the connexon. This hypothesis is supported by the findings that CaM co-localizes with GJs [85,162] and has a highly Ca2+-dependent affinity to all three intracellular connexin domains [163,164]. Furthermore, HC oligomerization and its opening can be prevented by the W7 CaM inhibitor [13,75,84,165] (Figure 2).
It is noteworthy to mention that the pathological G12R and N14K mutations of Cx26 both reduce the affinity of the NT to CaM [128,131]. In human and murine fibroblasts carrying the pathological G12S and S26L mutations of Cx32, an increased CaM-dependent protein kinase II (CaMKII) activity was linked to the CMT1X motor phenotype, mitotic instability, and HC dysfunction [166]. The defects were partially recovered by a CaMKII inhibitor (KN93), supporting the notion that a CaM-dependent pathway controls the HC gating by [Ca2+]c [13,75,167].
Overall, further investigation is needed to clarify the structural and chemical modifications of connexin HCs during opening by [Ca2+]e and [Ca2+]c variations. A more complete model accounting for Ca2+ and transmembrane voltage changes will undoubtedly improve our interpretation of the experimental results performed under physiological conditions. A model of this kind could open unprecedented opportunities for the development of therapeutic compounds that target specific HC dysfunctions. Recently, novel classes of molecules have been developed that directly interact with connexin HCs, including connexin mimetic peptides [168], anti-connexin antibodies [169], and aminoglycosides without antibiotic activity [170]. Although the mechanism of action for some of these molecules is not yet fully understood, their increased selectivity for specific connexin isoforms, combined with their reduced toxicity, makes them promising candidates for clinical application.

Author Contributions

E.B. and D.L.-P. collected the relevant information, drafted the figures and tables, and critically revised the manuscript; M.B. provided conceptualization of the text and figures, and drafted the manuscript. E.B. and D.L.-P. contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the University of Padua (grant BORT_BIRD23_01).

Acknowledgments

Images were created with BioRender.com (accessed on 18 April 2024).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Delmar, M.; Laird, D.W.; Naus, C.C.; Nielsen, M.S.; Verselis, V.K.; White, T.W. Connexins and Disease. Cold Spring Harb. Perspect. Biol. 2018, 10, a029348. [Google Scholar] [CrossRef] [PubMed]
  2. Epifantseva, I.; Shaw, R.M. Intracellular trafficking pathways of Cx43 gap junction channels. Biochim. Biophys. Acta Biomembr. 2018, 1860, 40–47. [Google Scholar] [CrossRef] [PubMed]
  3. Retamal, M.A.; Cortés, C.J.; Reuss, L.; Bennett, M.V.L.; Sáez, J.C. S-nitrosylation and permeation through connexin 43 hemichannels in astrocytes: Induction by oxidant stress and reversal by reducing agents. Proc. Natl. Acad. Sci. USA 2006, 103, 4475–4480. [Google Scholar] [CrossRef] [PubMed]
  4. Sanchez, H.A.; Orellana, J.A.; Verselis, V.K.; Saez, J.C. Metabolic inhibition increases activity of connexin-32 hemichannels permeable to Ca2+ in transfected HeLa cells. Am. J. Physiol. Cell Physiol. 2009, 297, C665–C678. [Google Scholar] [CrossRef] [PubMed]
  5. Harris, A.L. Emerging issues of connexin channels: Biophysics fills the gap. Q. Rev. Biophys. 2001, 34, 325–472. [Google Scholar] [CrossRef] [PubMed]
  6. Alexander, D.B.; Goldberg, G.S. Transfer of biologically important molecules between cells through gap junction channels. Curr. Med. Chem. 2003, 10, 2045–2058. [Google Scholar] [CrossRef] [PubMed]
  7. Saez, J.C.; Schalper, K.A.; Retamal, M.A.; Orellana, J.A.; Shoji, K.F.; Bennett, M.V. Cell membrane permeabilization via connexin hemichannels in living and dying cells. Exp. Cell Res. 2010, 316, 2377–2389. [Google Scholar] [CrossRef] [PubMed]
  8. Anselmi, F.; Hernandez, V.H.; Crispino, G.; Seydel, A.; Ortolano, S.; Roper, S.D.; Kessaris, N.; Richardson, W.; Rickheit, G.; Filippov, M.A.; et al. ATP release through connexin hemichannels and gap junction transfer of second messengers propagate Ca2+ signals across the inner ear. Proc. Natl. Acad. Sci. USA 2008, 105, 18770–18775. [Google Scholar] [CrossRef] [PubMed]
  9. Tran Van Nhieu, G.; Clair, C.; Bruzzone, R.; Mesnil, M.; Sansonetti, P.; Combettes, L. Connexin-dependent inter-cellular communication increases invasion and dissemination of Shigella in epithelial cells. Nat. Cell Biol. 2003, 5, 720–726. [Google Scholar] [CrossRef]
  10. Majumder, P.; Crispino, G.; Rodriguez, L.; Ciubotaru, C.D.; Anselmi, F.; Piazza, V.; Bortolozzi, M.; Mammano, F. ATP-mediated cell–cell signaling in the organ of Corti: The role of connexin channels. Purinergic Signal. 2010, 6, 167–187. [Google Scholar] [CrossRef]
  11. Sipos, A.; Vargas, S.L.; Toma, I.; Hanner, F.; Willecke, K.; Peti-Peterdi, J. Connexin 30 deficiency impairs renal tubular ATP release and pressure natriuresis. J. Am. Soc. Nephrol. 2009, 20, 1724–1732. [Google Scholar] [CrossRef] [PubMed]
  12. Liang, W.G.; Su, C.C.; Nian, J.H.; Chiang, A.S.; Li, S.Y.; Yang, J.J. Human connexin30.2/31.3 (GJC3) does not form functional gap junction channels but causes enhanced ATP release in HeLa cells. Cell Biochem. Biophys. 2011, 61, 189–197. [Google Scholar] [CrossRef] [PubMed]
  13. De Vuyst, E.; Decrock, E.; Cabooter, L.; Dubyak, G.R.; Naus, C.C.; Evans, W.H.; Leybaert, L. Intracellular calcium changes trigger connexin 32 hemichannel opening. EMBO J. 2006, 25, 34–44. [Google Scholar] [CrossRef]
  14. Schock, S.C.; Leblanc, D.; Hakim, A.M.; Thompson, C.S. ATP release by way of connexin 36 hemichannels mediates ischemic tolerance in vitro. Biochem. Biophys. Res. Commun. 2008, 368, 138–144. [Google Scholar] [CrossRef]
  15. Bahima, L.; Aleu, J.; Elias, M.; Martin-Satue, M.; Muhaisen, A.; Blasi, J.; Marsal, J.; Solsona, C. Endogenous hemichannels play a role in the release of ATP from Xenopus oocytes. J. Cell. Physiol. 2006, 206, 95–102. [Google Scholar] [CrossRef] [PubMed]
  16. Toma, I.; Bansal, E.; Meer, E.J.; Kang, J.J.; Vargas, S.L.; Peti-Peterdi, J. Connexin 40 and ATP-dependent intercellular calcium wave in renal glomerular endothelial cells. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008, 294, R1769–R1776. [Google Scholar] [CrossRef] [PubMed]
  17. Cotrina, M.L.; Lin, J.H.; Alves-Rodrigues, A.; Liu, S.; Li, J.; Azmi-Ghadimi, H.; Kang, J.; Naus, C.C.; Nedergaard, M. Connexins regulate calcium signaling by controlling ATP release. Proc. Natl. Acad. Sci. USA 1998, 95, 15735–15740. [Google Scholar] [CrossRef] [PubMed]
  18. Kang, J.; Kang, N.; Lovatt, D.; Torres, A.; Zhao, Z.; Lin, J.; Nedergaard, M. Connexin 43 hemichannels are permeable to ATP. J. Neurosci. 2008, 28, 4702–4711. [Google Scholar] [CrossRef] [PubMed]
  19. Lohman, A.W.; Billaud, M.; Isakson, B.E. Mechanisms of ATP release and signalling in the blood vessel wall. Cardiovasc. Res. 2012, 95, 269–280. [Google Scholar] [CrossRef]
  20. Stout, C.E.; Costantin, J.L.; Naus, C.C.; Charles, A.C. Intercellular calcium signaling in astrocytes via ATP release through connexin hemichannels. J. Biol. Chem. 2002, 277, 10482–10488. [Google Scholar] [CrossRef]
  21. Gomes, P.; Srinivas, S.P.; Van Driessche, W.; Vereecke, J.; Himpens, B. ATP release through connexin hemichannels in corneal endothelial cells. Investig. Ophthalmol. Vis. Sci. 2005, 46, 1208–1218. [Google Scholar] [CrossRef] [PubMed]
  22. Leybaert, L.; Sanderson, M.J. Intercellular Ca(2+) waves: Mechanisms and function. Physiol. Rev. 2012, 92, 1359–1392. [Google Scholar] [CrossRef] [PubMed]
  23. Valiunas, V. Cyclic nucleotide permeability through unopposed connexin hemichannels. Front. Pharmacol. 2013, 4, 75. [Google Scholar] [CrossRef] [PubMed]
  24. Ye, Z.C.; Wyeth, M.S.; Baltan-Tekkok, S.; Ransom, B.R. Functional hemichannels in astrocytes: A novel mechanism of glutamate release. J. Neurosci. 2003, 23, 3588–3596. [Google Scholar] [CrossRef] [PubMed]
  25. Takeuchi, H.; Jin, S.; Wang, J.; Zhang, G.; Kawanokuchi, J.; Kuno, R.; Sonobe, Y.; Mizuno, T.; Suzumura, A. Tumor necrosis factor-alpha induces neurotoxicity via glutamate release from hemichannels of activated microglia in an autocrine manner. J. Biol. Chem. 2006, 281, 21362–21368. [Google Scholar] [CrossRef] [PubMed]
  26. Bruzzone, S.; Guida, L.; Zocchi, E.; Franco, L.; De Flora, A. Connexin 43 hemi channels mediate Ca2+-regulated transmembrane NAD+ fluxes in intact cells. FASEB J. 2001, 15, 10–12. [Google Scholar] [CrossRef] [PubMed]
  27. Zocchi, E.; Podesta, M.; Pitto, A.; Usai, C.; Bruzzone, S.; Franco, L.; Guida, L.; Bacigalupo, A.; De Flora, A. Paracrinally stimulated expansion of early human hemopoietic progenitors by stroma-generated cyclic ADP-ribose. FASEB J. 2001, 15, 1610–1612. [Google Scholar] [CrossRef] [PubMed]
  28. Rana, S.; Dringen, R. Gap junction hemichannel-mediated release of glutathione from cultured rat astrocytes. Neurosci. Lett. 2007, 415, 45–48. [Google Scholar] [CrossRef]
  29. Linsambarth, S.; Carvajal, F.J.; Moraga-Amaro, R.; Mendez, L.; Tamburini, G.; Jimenez, I.; Verdugo, D.A.; Gomez, G.I.; Jury, N.; Martinez, P.; et al. Astroglial gliotransmitters released via Cx43 hemichannels regulate NMDAR-dependent transmission and short-term fear memory in the basolateral amygdala. FASEB J. 2022, 36, e22134. [Google Scholar] [CrossRef]
  30. Cherian, P.P.; Siller-Jackson, A.J.; Gu, S.; Wang, X.; Bonewald, L.F.; Sprague, E.; Jiang, J.X. Mechanical strain opens connexin 43 hemichannels in osteocytes: A novel mechanism for the release of prostaglandin. Mol. Biol. Cell 2005, 16, 3100–3106. [Google Scholar] [CrossRef]
  31. Burra, S.; Jiang, J.X. Connexin 43 hemichannel opening associated with Prostaglandin E(2) release is adaptively regulated by mechanical stimulation. Commun. Integr. Biol. 2009, 2, 239–240. [Google Scholar] [CrossRef] [PubMed]
  32. Song, D.; Liu, X.; Liu, R.; Yang, L.; Zuo, J.; Liu, W. Connexin 43 hemichannel regulates H9c2 cell proliferation by modulating intracellular ATP and [Ca2+]. Acta Biochim. Biophys. Sin. 2010, 42, 472–482. [Google Scholar] [CrossRef] [PubMed]
  33. Cotrina, M.L.; Lin, J.H.; Nedergaard, M. Adhesive properties of connexin hemichannels. Glia 2008, 56, 1791–1798. [Google Scholar] [CrossRef] [PubMed]
  34. Kamermans, M.; Fahrenfort, I.; Schultz, K.; Janssen-Bienhold, U.; Sjoerdsma, T.; Weiler, R. Hemichannel-mediated inhibition in the outer retina. Science 2001, 292, 1178–1180. [Google Scholar] [CrossRef] [PubMed]
  35. Mathias, R.T.; White, T.W.; Gong, X. Lens gap junctions in growth, differentiation, and homeostasis. Physiol. Rev. 2010, 90, 179–206. [Google Scholar] [CrossRef] [PubMed]
  36. Beyer, E.C.; Berthoud, V.M. Connexin hemichannels in the lens. Front. Physiol. 2014, 5, 20. [Google Scholar] [CrossRef] [PubMed]
  37. Ceriani, F.; Pozzan, T.; Mammano, F. Critical role of ATP-induced ATP release for Ca2+ signaling in nonsensory cell networks of the developing cochlea. Proc. Natl. Acad. Sci. USA 2016, 113, E7194–E7201. [Google Scholar] [CrossRef] [PubMed]
  38. Tritsch, N.X.; Bergles, D.E. Developmental regulation of spontaneous activity in the Mammalian cochlea. J. Neurosci. 2010, 30, 1539–1550. [Google Scholar] [CrossRef] [PubMed]
  39. Kar, R.; Riquelme, M.A.; Werner, S.; Jiang, J.X. Connexin 43 channels protect osteocytes against oxidative stress-induced cell death. J. Bone Miner. Res. 2013, 28, 1611–1621. [Google Scholar] [CrossRef]
  40. Lecanda, F.; Warlow, P.M.; Sheikh, S.; Furlan, F.; Steinberg, T.H.; Civitelli, R. Connexin43 deficiency causes delayed ossification, craniofacial abnormalities, and osteoblast dysfunction. J. Cell Biol. 2000, 151, 931–944. [Google Scholar] [CrossRef]
  41. Riquelme, M.A.; Jiang, J.X. Elevated Intracellular Ca2+ Signals by Oxidative Stress Activate Connexin 43 Hemichannels in Osteocytes. Bone Res. 2013, 1, 355–361. [Google Scholar] [CrossRef] [PubMed]
  42. Zeng, Y.; Riquelme, M.A.; Hua, R.; Zhang, J.; Acosta, F.M.; Gu, S.; Jiang, J.X. Mechanosensitive piezo1 calcium channel activates connexin 43 hemichannels through PI3K signaling pathway in bone. Cell Biosci. 2022, 12, 191. [Google Scholar] [CrossRef] [PubMed]
  43. Quist, A.P.; Rhee, S.K.; Lin, H.; Lal, R. Physiological role of gap-junctional hemichannels. Extracellular calcium-dependent isosmotic volume regulation. J. Cell Biol. 2000, 148, 1063–1074. [Google Scholar] [CrossRef] [PubMed]
  44. De Bock, M.; Culot, M.; Wang, N.; Bol, M.; Decrock, E.; De Vuyst, E.; da Costa, A.; Dauwe, I.; Vinken, M.; Simon, A.M.; et al. Connexin channels provide a target to manipulate brain endothelial calcium dynamics and blood-brain barrier permeability. J. Cereb. Blood Flow. Metab. 2011, 31, 1942–1957. [Google Scholar] [CrossRef] [PubMed]
  45. Bosch, M.; Kielian, T. Hemichannels in neurodegenerative diseases: Is there a link to pathology? Front. Cell. Neurosci. 2014, 8, 242. [Google Scholar] [CrossRef] [PubMed]
  46. Crunelli, V.; Carmignoto, G.; Steinhauser, C. Novel astrocyte targets: New avenues for the therapeutic treatment of epilepsy. Neuroscientist 2015, 21, 62–83. [Google Scholar] [CrossRef] [PubMed]
  47. Retamal, M.A.; Reyes, E.P.; Garcia, I.E.; Pinto, B.; Martinez, A.D.; Gonzalez, C. Diseases associated with leaky hemichannels. Front. Cell. Neurosci. 2015, 9, 267. [Google Scholar] [CrossRef] [PubMed]
  48. Orellana, J.A.; Stehberg, J. Hemichannels: New roles in astroglial function. Front. Physiol. 2014, 5, 193. [Google Scholar] [CrossRef] [PubMed]
  49. Stehberg, J.; Moraga-Amaro, R.; Salazar, C.; Becerra, A.; Echeverria, C.; Orellana, J.A.; Bultynck, G.; Ponsaerts, R.; Leybaert, L.; Simon, F.; et al. Release of gliotransmitters through astroglial connexin 43 hemichannels is necessary for fear memory consolidation in the basolateral amygdala. FASEB J. 2012, 26, 3649–3657. [Google Scholar] [CrossRef]
  50. Meunier, C.; Wang, N.; Yi, C.; Dallerac, G.; Ezan, P.; Koulakoff, A.; Leybaert, L.; Giaume, C. Contribution of Astroglial Cx43 Hemichannels to the Modulation of Glutamatergic Currents by D-Serine in the Mouse Prefrontal Cortex. J. Neurosci. 2017, 37, 9064–9075. [Google Scholar] [CrossRef]
  51. Bejarano, E.; Yuste, A.; Patel, B.; Stout, R.F., Jr.; Spray, D.C.; Cuervo, A.M. Connexins modulate autophagosome biogenesis. Nat. Cell Biol. 2014, 16, 401–414. [Google Scholar] [CrossRef]
  52. Schalper, K.A.; Carvajal-Hausdorf, D.; Oyarzo, M.P. Possible role of hemichannels in cancer. Front. Physiol. 2014, 5, 237. [Google Scholar] [CrossRef] [PubMed]
  53. Gadicherla, A.K.; Wang, N.; Bulic, M.; Agullo-Pascual, E.; Lissoni, A.; De Smet, M.; Delmar, M.; Bultynck, G.; Krysko, D.V.; Camara, A.; et al. Mitochondrial Cx43 hemichannels contribute to mitochondrial calcium entry and cell death in the heart. Basic Res. Cardiol. 2017, 112, 27. [Google Scholar] [CrossRef] [PubMed]
  54. Miro-Casas, E.; Ruiz-Meana, M.; Agullo, E.; Stahlhofen, S.; Rodriguez-Sinovas, A.; Cabestrero, A.; Jorge, I.; Torre, I.; Vazquez, J.; Boengler, K.; et al. Connexin43 in cardiomyocyte mitochondria contributes to mitochondrial potassium uptake. Cardiovasc. Res. 2009, 83, 747–756. [Google Scholar] [CrossRef]
  55. Bukauskas, F.F.; Kreuzberg, M.M.; Rackauskas, M.; Bukauskiene, A.; Bennett, M.V.; Verselis, V.K.; Willecke, K. Properties of mouse connexin 30.2 and human connexin 31.9 hemichannels: Implications for atrioventricular conduction in the heart. Proc. Natl. Acad. Sci. USA 2006, 103, 9726–9731. [Google Scholar] [CrossRef] [PubMed]
  56. De Smet, M.A.J.; Lissoni, A.; Nezlobinsky, T.; Wang, N.; Dries, E.; Pérez-Hernández, M.; Lin, X.; Amoni, M.; Vervliet, T.; Witschas, K.; et al. Cx43 hemichannel microdomain signaling at the intercalated disc enhances cardiac excitability. J. Clin. Investig. 2021, 131, e137752. [Google Scholar] [CrossRef]
  57. Bol, M.; Wang, N.; De Bock, M.; Wacquier, B.; Decrock, E.; Gadicherla, A.; Decaluwe, K.; Vanheel, B.; van Rijen, H.V.; Krysko, D.V.; et al. At the cross-point of connexins, calcium, and ATP: Blocking hemichannels inhibits vasoconstriction of rat small mesenteric arteries. Cardiovasc. Res. 2017, 113, 195–206. [Google Scholar] [CrossRef]
  58. Kuang, Y.; Zorzi, V.; Buratto, D.; Ziraldo, G.; Mazzarda, F.; Peres, C.; Nardin, C.; Salvatore, A.M.; Chiani, F.; Scavizzi, F.; et al. A potent antagonist antibody targeting connexin hemichannels alleviates Clouston syndrome symptoms in mutant mice. EBioMedicine 2020, 57, 102825. [Google Scholar] [CrossRef]
  59. Contreras, J.E.; Saez, J.C.; Bukauskas, F.F.; Bennett, M.V. Gating and regulation of connexin 43 (Cx43) hemichannels. Proc. Natl. Acad. Sci. USA 2003, 100, 11388–11393. [Google Scholar] [CrossRef]
  60. Valiunas, V.; Weingart, R. Electrical properties of gap junction hemichannels identified in transfected HeLa cells. Pflugers Arch. 2000, 440, 366–379. [Google Scholar] [CrossRef]
  61. Retamal, M.A. Connexin and Pannexin hemichannels are regulated by redox potential. Front. Physiol. 2014, 5, 80. [Google Scholar] [CrossRef] [PubMed]
  62. De Vuyst, E.; Decrock, E.; De Bock, M.; Yamasaki, H.; Naus, C.C.; Evans, W.H.; Leybaert, L. Connexin hemichannels and gap junction channels are differentially influenced by lipopolysaccharide and basic fibroblast growth factor. Mol. Biol. Cell 2007, 18, 34–46. [Google Scholar] [CrossRef] [PubMed]
  63. Li, H.; Liu, T.F.; Lazrak, A.; Peracchia, C.; Goldberg, G.S.; Lampe, P.D.; Johnson, R.G. Properties and regulation of gap junctional hemichannels in the plasma membranes of cultured cells. J. Cell Biol. 1996, 134, 1019–1030. [Google Scholar] [CrossRef] [PubMed]
  64. Bao, L.; Sachs, F.; Dahl, G. Connexins are mechanosensitive. Am. J. Physiol. Cell Physiol. 2004, 287, C1389–C1395. [Google Scholar] [CrossRef] [PubMed]
  65. Svenningsen, P.; Burford, J.L.; Peti-Peterdi, J. ATP releasing connexin 30 hemichannels mediate flow-induced calcium signaling in the collecting duct. Front. Physiol. 2013, 4, 292. [Google Scholar] [CrossRef] [PubMed]
  66. Bevans, C.G.; Harris, A.L. Regulation of connexin channels by pH. Direct action of the protonated form of taurine and other aminosulfonates. J. Biol. Chem. 1999, 274, 3711–3719. [Google Scholar] [CrossRef] [PubMed]
  67. Huckstepp, R.T.; Eason, R.; Sachdev, A.; Dale, N. CO2-dependent opening of connexin 26 and related beta connexins. J. Physiol. 2010, 588 Pt 20, 3921–3931. [Google Scholar] [CrossRef] [PubMed]
  68. Trexler, E.B.; Bukauskas, F.F.; Bennett, M.V.; Bargiello, T.A.; Verselis, V.K. Rapid and direct effects of pH on connexins revealed by the connexin46 hemichannel preparation. J. Gen. Physiol. 1999, 113, 721–742. [Google Scholar] [CrossRef] [PubMed]
  69. Delmar, M.; Coombs, W.; Sorgen, P.; Duffy, H.S.; Taffet, S.M. Structural bases for the chemical regulation of Connexin43 channels. Cardiovasc. Res. 2004, 62, 268–275. [Google Scholar] [CrossRef]
  70. Srinivas, M.; Calderon, D.P.; Kronengold, J.; Verselis, V.K. Regulation of connexin hemichannels by monovalent cations. J. Gen. Physiol. 2006, 127, 67–75. [Google Scholar] [CrossRef]
  71. Ebihara, L.; Liu, X.; Pal, J.D. Effect of external magnesium and calcium on human connexin46 hemichannels. Biophys. J. 2003, 84, 277–286. [Google Scholar] [CrossRef] [PubMed]
  72. Gomez-Hernandez, J.M.; de Miguel, M.; Larrosa, B.; Gonzalez, D.; Barrio, L.C. Molecular basis of calcium regulation in connexin-32 hemichannels. Proc. Natl. Acad. Sci. USA 2003, 100, 16030–16035. [Google Scholar] [CrossRef] [PubMed]
  73. Lopez, W.; Ramachandran, J.; Alsamarah, A.; Luo, Y.; Harris, A.L.; Contreras, J.E. Mechanism of gating by calcium in connexin hemichannels. Proc. Natl. Acad. Sci. USA 2016, 113, E7986–E7995. [Google Scholar] [CrossRef] [PubMed]
  74. Carrer, A.; Leparulo, A.; Crispino, G.; Ciubotaru, C.D.; Marin, O.; Zonta, F.; Bortolozzi, M. Cx32 hemichannel opening by cytosolic Ca2+ is inhibited by the R220X mutation that causes Charcot-Marie-Tooth disease. Hum. Mol. Genet. 2018, 27, 80–94. [Google Scholar] [CrossRef] [PubMed]
  75. De Vuyst, E.; Wang, N.; Decrock, E.; De Bock, M.; Vinken, M.; Van Moorhem, M.; Lai, C.; Culot, M.; Rogiers, V.; Cecchelli, R.; et al. Ca2+ regulation of connexin 43 hemichannels in C6 glioma and glial cells. Cell Calcium 2009, 46, 176–187. [Google Scholar] [CrossRef] [PubMed]
  76. Breitwieser, G.E. Extracellular calcium as an integrator of tissue function. Int. J. Biochem. Cell Biol. 2008, 40, 1467–1480. [Google Scholar] [CrossRef]
  77. Bosher, S.K.; Warren, R.L. Very low calcium content of cochlear endolymph, an extracellular fluid. Nature 1978, 273, 377–378. [Google Scholar] [CrossRef]
  78. Lopez, W.; Gonzalez, J.; Liu, Y.; Harris, A.L.; Contreras, J.E. Insights on the mechanisms of Ca2+ regulation of connexin26 hemichannels revealed by human pathogenic mutations (D50N/Y). J. Gen. Physiol. 2013, 142, 23–35. [Google Scholar] [CrossRef] [PubMed]
  79. Puljung, M.C.; Berthoud, V.M.; Beyer, E.C.; Hanck, D.A. Polyvalent Cations Constitute the Voltage Gating Particle in Human Connexin37 Hemichannels. J. Gen. Physiol. 2004, 124, 587–603. [Google Scholar] [CrossRef]
  80. Vargas, A.A.; Cisterna, B.A.; Saavedra-Leiva, F.; Urrutia, C.; Cea, L.A.; Vielma, A.H.; Gutierrez-Maldonado, S.E.; Martin, A.J.; Pareja-Barrueto, C.; Escalona, Y.; et al. On Biophysical Properties and Sensitivity to Gap Junction Blockers of Connexin 39 Hemichannels Expressed in HeLa Cells. Front. Physiol. 2017, 8, 38. [Google Scholar] [CrossRef]
  81. Allen, M.J.; Gemel, J.; Beyer, E.C.; Lal, R. Atomic force microscopy of Connexin40 gap junction hemichannels reveals calcium-dependent three-dimensional molecular topography and open-closed conformations of both the extracellular and cytoplasmic faces. J. Biol. Chem. 2011, 286, 22139–22146. [Google Scholar] [CrossRef] [PubMed]
  82. Bader, P.; Weingart, R.; Egger, M. Regulation of Cx45 hemichannels mediated by extracellular and intracellular calcium. Pflugers Arch. 2012, 464, 249–259. [Google Scholar] [CrossRef] [PubMed]
  83. Pfahnl, A.; Dahl, G. Gating of cx46 gap junction hemichannels by calcium and voltage. Pflügers Arch. 1999, 437, 345–353. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, X.; Zou, T.; Liu, Y.; Qi, Y. The gating effect of calmodulin and calcium on the connexin50 hemichannel. Biol. Chem. 2006, 387, 595–601. [Google Scholar] [CrossRef] [PubMed]
  85. Hu, Z.; Riquelme, M.A.; Wang, B.; Bugay, V.; Brenner, R.; Gu, S.; Jiang, J.X. Cataract-associated Connexin 46 Mutation Alters Its Interaction with Calmodulin and Function of Hemichannels. J. Biol. Chem. 2018, 293, 2573–2585. [Google Scholar] [CrossRef] [PubMed]
  86. Torres, A.; Wang, F.; Xu, Q.; Fujita, T.; Dobrowolski, R.; Willecke, K.; Takano, T.; Nedergaard, M. Extracellular Ca2+ Acts as a Mediator of Communication from Neurons to Glia. Science Signal. 2012, 5, ra8. [Google Scholar] [CrossRef] [PubMed]
  87. Ceriani, F.; Hendry, A.; Jeng, J.Y.; Johnson, S.L.; Stephani, F.; Olt, J.; Holley, M.C.; Mammano, F.; Engel, J.; Kros, C.J.; et al. Coordinated calcium signalling in cochlear sensory and non-sensory cells refines afferent innervation of outer hair cells. EMBO J. 2019, 38, e99839. [Google Scholar] [CrossRef] [PubMed]
  88. Lee, S.E.; Lee, S.H. Skin Barrier and Calcium. Ann. Dermatol. 2018, 30, 265–275. [Google Scholar] [CrossRef] [PubMed]
  89. Muller, D.J.; Hand, G.M.; Engel, A.; Sosinsky, G.E. Conformational changes in surface structures of isolated connexin 26 gap junctions. EMBO J. 2002, 21, 3598–3607. [Google Scholar] [CrossRef]
  90. Ripps, H.; Qian, H.; Zakevicius, J. Properties of connexin26 hemichannels expressed in Xenopus oocytes. Cell. Mol. Neurobiol. 2004, 24, 647–665. [Google Scholar] [CrossRef]
  91. Zonta, F.; Mammano, F.; Torsello, M.; Fortunati, N.; Orian, L.; Polimeno, A. Role of gamma carboxylated Glu47 in connexin 26 hemichannel regulation by extracellular Ca2+: Insight from a local quantum chemistry study. Biochem. Biophys. Res. Commun. 2014, 445, 10–15. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, Y.; Hao, H. Conserved glycine at position 45 of major cochlear connexins constitutes a vital component of the Ca2+ sensor for gating of gap junction hemichannels. Biochem. Biophys. Res. Commun. 2013, 436, 424–429. [Google Scholar] [CrossRef] [PubMed]
  93. Pinto, B.I.; Pupo, A.; Garcia, I.E.; Mena-Ulecia, K.; Martinez, A.D.; Latorre, R.; Gonzalez, C. Calcium binding and voltage gating in Cx46 hemichannels. Sci. Rep. 2017, 7, 15851. [Google Scholar] [CrossRef] [PubMed]
  94. Zonta, F.; Polles, G.; Zanotti, G.; Mammano, F. Permeation pathway of homomeric connexin 26 and connexin 30 channels investigated by molecular dynamics. J. Biomol. Struct. Dyn. 2012, 29, 985–998. [Google Scholar] [CrossRef] [PubMed]
  95. Fasciani, I.; Temperan, A.; Perez-Atencio, L.F.; Escudero, A.; Martinez-Montero, P.; Molano, J.; Gomez-Hernandez, J.M.; Paino, C.L.; Gonzalez-Nieto, D.; Barrio, L.C. Regulation of connexin hemichannel activity by membrane potential and the extracellular calcium in health and disease. Neuropharmacology 2013, 75, 479–490. [Google Scholar] [CrossRef]
  96. Choi, E.J.; Palacios-Prado, N.; Saez, J.C.; Lee, J. Identification of Cx45 as a Major Component of GJs in HeLa Cells. Biomolecules 2020, 10, 1389. [Google Scholar]
  97. Verselis, V.K.; Srinivas, M. Divalent cations regulate connexin hemichannels by modulating intrinsic voltage-dependent gating. J. Gen. Physiol. 2008, 132, 315–327. [Google Scholar] [CrossRef] [PubMed]
  98. Beahm, D.L.; Hall, J.E. Hemichannel and junctional properties of connexin 50. Biophys. J. 2002, 82, 2016–2031. [Google Scholar] [CrossRef] [PubMed]
  99. Bootman, M.D.; Bultynck, G. Fundamentals of Cellular Calcium Signaling: A Primer. Cold Spring Harb. Perspect. Biol. 2020, 12, a038802. [Google Scholar] [CrossRef]
  100. Bagur, R.; Hajnoczky, G. Intracellular Ca2+ Sensing: Its Role in Calcium Homeostasis and Signaling. Mol. Cell 2017, 66, 780–788. [Google Scholar] [CrossRef]
  101. Lurtz, M.M.; Louis, C.F. Intracellular calcium regulation of connexin43. Am. J. Physiol. Cell Physiol. 2007, 293, C1806–C1813. [Google Scholar] [CrossRef] [PubMed]
  102. Figueroa, V.; Saez, P.J.; Salas, J.D.; Salas, D.; Jara, O.; Martinez, A.D.; Saez, J.C.; Retamal, M.A. Linoleic acid induces opening of connexin26 hemichannels through a PI3K/Akt/Ca2+-dependent pathway. Biochim. Biophys. Acta 2013, 1828, 1169–1179. [Google Scholar] [CrossRef] [PubMed]
  103. Schalper, K.A.; Palacios-Prado, N.; Retamal, M.A.; Shoji, K.F.; Martinez, A.D.; Saez, J.C. Connexin hemichannel composition determines the FGF-1-induced membrane permeability and free [Ca2+]i responses. Mol. Biol. Cell 2008, 19, 3501–3513. [Google Scholar] [CrossRef] [PubMed]
  104. Lagos-Cabre, R.; Brenet, M.; Diaz, J.; Perez, R.D.; Perez, L.A.; Herrera-Molina, R.; Quest, A.F.G.; Leyton, L. Intracellular Ca2+ Increases and Connexin 43 Hemichannel Opening Are Necessary but Not Sufficient for Thy-1-Induced Astrocyte Migration. Int. J. Mol. Sci. 2018, 19, 2179. [Google Scholar] [CrossRef] [PubMed]
  105. Pearson, R.A.; Dale, N.; Llaudet, E.; Mobbs, P. ATP released via gap junction hemichannels from the pigment epithelium regulates neural retinal progenitor proliferation. Neuron 2005, 46, 731–744. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, N.; De Vuyst, E.; Ponsaerts, R.; Boengler, K.; Palacios-Prado, N.; Wauman, J.; Lai, C.P.; De Bock, M.; Decrock, E.; Bol, M.; et al. Selective inhibition of Cx43 hemichannels by Gap19 and its impact on myocardial ischemia/reperfusion injury. Basic Res. Cardiol. 2013, 108, 309. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, N.; De Bock, M.; Antoons, G.; Gadicherla, A.K.; Bol, M.; Decrock, E.; Evans, W.H.; Sipido, K.R.; Bukauskas, F.F.; Leybaert, L. Connexin mimetic peptides inhibit Cx43 hemichannel opening triggered by voltage and intracellular Ca2+ elevation. Basic. Res. Cardiol. 2012, 107, 304. [Google Scholar] [CrossRef] [PubMed]
  108. Braet, K.; Vandamme, W.; Martin, P.E.; Evans, W.H.; Leybaert, L. Photoliberating inositol-1,4,5-trisphosphate triggers ATP release that is blocked by the connexin mimetic peptide gap 26. Cell Calcium 2003, 33, 37–48. [Google Scholar] [CrossRef] [PubMed]
  109. Arcuino, G.; Lin, J.H.; Takano, T.; Liu, C.; Jiang, L.; Gao, Q.; Kang, J.; Nedergaard, M. Intercellular calcium signaling mediated by point-source burst release of ATP. Proc. Natl. Acad. Sci. USA 2002, 99, 9840–9845. [Google Scholar] [CrossRef]
  110. Van Campenhout, R.; Gomes, A.R.; De Groof, T.W.M.; Muyldermans, S.; Devoogdt, N.; Vinken, M. Mechanisms Underlying Connexin Hemichannel Activation in Disease. Int. J. Mol. Sci. 2021, 22, 3503. [Google Scholar] [CrossRef]
  111. Guo, A.; Zhang, H.; Li, H.; Chiu, A.; Garcia-Rodriguez, C.; Lagos, C.F.; Saez, J.C.; Lau, C.G. Inhibition of connexin hemichannels alleviates neuroinflammation and hyperexcitability in temporal lobe epilepsy. Proc. Natl. Acad. Sci. USA 2022, 119, e2213162119. [Google Scholar] [CrossRef] [PubMed]
  112. Gonzalez-Jamett, A.; Vasquez, W.; Cifuentes-Riveros, G.; Martinez-Pando, R.; Saez, J.C.; Cardenas, A.M. Oxidative Stress, Inflammation and Connexin Hemichannels in Muscular Dystrophies. Biomedicines 2022, 10, 507. [Google Scholar] [CrossRef] [PubMed]
  113. Verselis, V.K. Connexin hemichannels and cochlear function. Neurosci. Lett. 2019, 695, 40–45. [Google Scholar] [CrossRef] [PubMed]
  114. Garcia, I.E.; Bosen, F.; Mujica, P.; Pupo, A.; Flores-Munoz, C.; Jara, O.; Gonzalez, C.; Willecke, K.; Martinez, A.D. From Hyperactive Connexin26 Hemichannels to Impairments in Epidermal Calcium Gradient and Permeability Barrier in the Keratitis-Ichthyosis-Deafness Syndrome. J. Investig. Dermatol. 2016, 136, 574–583. [Google Scholar] [CrossRef] [PubMed]
  115. Bortolozzi, M. What’s the Function of Connexin 32 in the Peripheral Nervous System? Front. Mol. Neurosci. 2018, 11, 227. [Google Scholar] [CrossRef] [PubMed]
  116. Kelly, J.J.; Esseltine, J.L.; Shao, Q.; Jabs, E.W.; Sampson, J.; Auranen, M.; Bai, D.; Laird, D.W. Specific functional pathologies of Cx43 mutations associated with oculodentodigital dysplasia. Mol. Biol. Cell 2016, 27, 2172–2185. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, H.; Cao, X.; Lin, Z.; Lee, M.; Jia, X.; Ren, Y.; Dai, L.; Guan, L.; Zhang, J.; Lin, X.; et al. Exome sequencing reveals mutation in GJA1 as a cause of keratoderma-hypotrichosis-leukonychia totalis syndrome. Hum. Mol. Genet. 2015, 24, 6564. [Google Scholar] [CrossRef]
  118. Lucaciu, S.A.; Figliuzzi, R.; Neumann, R.; Nazarali, S.; Del Sordo, L.; Leighton, S.E.; Hauser, A.; Shao, Q.; Johnston, D.; Bai, D.; et al. GJB4 variants linked to skin disease exhibit a trafficking deficiency en route to gap junction formation that can be restored by co-expression of select connexins. Front. Cell Dev. Biol. 2023, 11, 1073805. [Google Scholar] [CrossRef] [PubMed]
  119. Beyer, E.C.; Ebihara, L.; Berthoud, V.M. Connexin mutants and cataracts. Front. Pharmacol. 2013, 4, 43. [Google Scholar] [CrossRef]
  120. Prieto-Villalobos, J.; Alvear, T.F.; Liberona, A.; Lucero, C.M.; Martinez-Araya, C.J.; Balmazabal, J.; Inostroza, C.A.; Ramirez, G.; Gomez, G.I.; Orellana, J.A. Astroglial Hemichannels and Pannexons: The Hidden Link between Maternal Inflammation and Neurological Disorders. Int. J. Mol. Sci. 2021, 22, 9503. [Google Scholar] [CrossRef]
  121. Lucero, C.M.; Prieto-Villalobos, J.; Marambio-Ruiz, L.; Balmazabal, J.; Alvear, T.F.; Vega, M.; Barra, P.; Retamal, M.A.; Orellana, J.A.; Gomez, G.I. Hypertensive Nephropathy: Unveiling the Possible Involvement of Hemichannels and Pannexons. Int. J. Mol. Sci. 2022, 23, 15936. [Google Scholar] [CrossRef]
  122. Dobrowolski, R.; Sommershof, A.; Willecke, K. Some oculodentodigital dysplasia-associated Cx43 mutations cause increased hemichannel activity in addition to deficient gap junction channels. J. Membr. Biol. 2007, 219, 9–17. [Google Scholar] [CrossRef]
  123. Lee, J.R.; Derosa, A.M.; White, T.W. Connexin mutations causing skin disease and deafness increase hemichannel activity and cell death when expressed in Xenopus oocytes. J. Investig. Dermatol. 2009, 129, 870–878. [Google Scholar] [CrossRef]
  124. Ebihara, L.; Korzyukov, Y.; Kothari, S.; Tong, J.J. Cx46 hemichannels contribute to the sodium leak conductance in lens fiber cells. Am. J. Physiol. Cell Physiol. 2014, 306, C506–C513. [Google Scholar] [CrossRef] [PubMed]
  125. Martinez, A.D.; Acuna, R.; Figueroa, V.; Maripillan, J.; Nicholson, B. Gap-junction channels dysfunction in deafness and hearing loss. Antioxid. Redox Signal. 2009, 11, 309–322. [Google Scholar] [CrossRef]
  126. Xu, J.; Nicholson, B.J. The role of connexins in ear and skin physiology-functional insights from disease-associated mutations. Biochim. Biophys. Acta 2013, 1828, 167–178. [Google Scholar] [CrossRef]
  127. Terrinoni, A.; Codispoti, A.; Serra, V.; Didona, B.; Bruno, E.; Nistico, R.; Giustizieri, M.; Alessandrini, M.; Campione, E.; Melino, G. Connexin 26 (GJB2) mutations, causing KID Syndrome, are associated with cell death due to calcium gating deregulation. Biochem. Biophys. Res. Commun. 2010, 394, 909–914. [Google Scholar] [CrossRef]
  128. Garcia, I.E.; Villanelo, F.; Contreras, G.F.; Pupo, A.; Pinto, B.I.; Contreras, J.E.; Perez-Acle, T.; Alvarez, O.; Latorre, R.; Martinez, A.D.; et al. The syndromic deafness mutation G12R impairs fast and slow gating in Cx26 hemichannels. J. Gen. Physiol. 2018, 150, 697–711. [Google Scholar] [CrossRef] [PubMed]
  129. Garcia, I.E.; Maripillan, J.; Jara, O.; Ceriani, R.; Palacios-Munoz, A.; Ramachandran, J.; Olivero, P.; Perez-Acle, T.; Gonzalez, C.; Saez, J.C.; et al. Keratitis-ichthyosis-deafness syndrome-associated Cx26 mutants produce nonfunctional gap junctions but hyperactive hemichannels when co-expressed with wild type Cx43. J. Investig. Dermatol. 2015, 135, 1338–1347. [Google Scholar] [CrossRef]
  130. Press, E.R.; Shao, Q.; Kelly, J.J.; Chin, K.; Alaga, A.; Laird, D.W. Induction of cell death and gain-of-function properties of connexin26 mutants predict severity of skin disorders and hearing loss. J. Biol. Chem. 2017, 292, 9721–9732. [Google Scholar] [CrossRef]
  131. Valdez Capuccino, J.M.; Chatterjee, P.; Garcia, I.E.; Botello-Smith, W.M.; Zhang, H.; Harris, A.L.; Luo, Y.; Contreras, J.E. The connexin26 human mutation N14K disrupts cytosolic intersubunit interactions and promotes channel opening. J. Gen. Physiol. 2019, 151, 328–341. [Google Scholar] [CrossRef]
  132. Aypek, H.; Bay, V.; Mese, G. Altered cellular localization and hemichannel activities of KID syndrome associated connexin26 I30N and D50Y mutations. BMC Cell Biol. 2016, 17, 5. [Google Scholar] [CrossRef]
  133. Sanchez, H.A.; Mese, G.; Srinivas, M.; White, T.W.; Verselis, V.K. Differentially altered Ca2+ regulation and Ca2+ permeability in Cx26 hemichannels formed by the A40V and G45E mutations that cause keratitis ichthyosis deafness syndrome. J. Gen. Physiol. 2010, 136, 47–62. [Google Scholar] [CrossRef]
  134. Mhaske, P.V.; Levit, N.A.; Li, L.; Wang, H.Z.; Lee, J.R.; Shuja, Z.; Brink, P.R.; White, T.W. The human Cx26-D50A and Cx26-A88V mutations causing keratitis-ichthyosis-deafness syndrome display increased hemichannel activity. Am. J. Physiol. Cell Physiol. 2013, 304, C1150–C1158. [Google Scholar] [CrossRef]
  135. Chen, Y.; Deng, Y.; Bao, X.; Reuss, L.; Altenberg, G.A. Mechanism of the defect in gap-junctional communication by expression of a connexin 26 mutant associated with dominant deafness. FASEB J. 2005, 19, 1516–1518. [Google Scholar] [CrossRef]
  136. Bosen, F.; Celli, A.; Crumrine, D.; vom Dorp, K.; Ebel, P.; Jastrow, H.; Dormann, P.; Winterhager, E.; Mauro, T.; Willecke, K. Altered epidermal lipid processing and calcium distribution in the KID syndrome mouse model Cx26S17F. FEBS Lett. 2015, 589, 1904–1910. [Google Scholar] [CrossRef]
  137. Lopez, W.; Liu, Y.; Harris, A.L.; Contreras, J.E. Divalent regulation and intersubunit interactions of human Connexin26 (Cx26) hemichannels. Channels 2014, 8, 1–4. [Google Scholar] [CrossRef]
  138. Abbott, A.C.; Garcia, I.E.; Villanelo, F.; Flores-Munoz, C.; Ceriani, R.; Maripillan, J.; Novoa-Molina, J.; Figueroa-Cares, C.; Perez-Acle, T.; Saez, J.C.; et al. Expression of KID syndromic mutation Cx26S17F produces hyperactive hemichannels in supporting cells of the organ of Corti. Front. Cell Dev. Biol. 2022, 10, 1071202. [Google Scholar] [CrossRef]
  139. Jara, O.; Acuna, R.; Garcia, I.E.; Maripillan, J.; Figueroa, V.; Saez, J.C.; Araya-Secchi, R.; Lagos, C.F.; Perez-Acle, T.; Berthoud, V.M.; et al. Critical role of the first transmembrane domain of Cx26 in regulating oligomerization and function. Mol. Biol. Cell 2012, 23, 3299–3311. [Google Scholar] [CrossRef]
  140. Stong, B.C.; Chang, Q.; Ahmad, S.; Lin, X. A novel mechanism for connexin 26 mutation linked deafness: Cell death caused by leaky gap junction hemichannels. Laryngoscope 2006, 116, 2205–2210. [Google Scholar] [CrossRef]
  141. Essenfelder, G.M.; Bruzzone, R.; Lamartine, J.; Charollais, A.; Blanchet-Bardon, C.; Barbe, M.T.; Meda, P.; Waksman, G. Connexin30 mutations responsible for hidrotic ectodermal dysplasia cause abnormal hemichannel activity. Hum. Mol. Genet. 2004, 13, 1703–1714. [Google Scholar] [CrossRef]
  142. Abrams, C.K.; Bennett, M.V.; Verselis, V.K.; Bargiello, T.A. Voltage opens unopposed gap junction hemichannels formed by a connexin 32 mutant associated with X-linked Charcot-Marie-Tooth disease. Proc. Natl. Acad. Sci. USA 2002, 99, 3980–3984. [Google Scholar] [CrossRef]
  143. Srinivas, M.; Jannace, T.F.; Cocozzelli, A.G.; Li, L.; Slavi, N.; Sellitto, C.; White, T.W. Connexin43 mutations linked to skin disease have augmented hemichannel activity. Sci. Rep. 2019, 9, 19. [Google Scholar] [CrossRef]
  144. Lai, A.; Le, D.N.; Paznekas, W.A.; Gifford, W.D.; Jabs, E.W.; Charles, A.C. Oculodentodigital dysplasia connexin43 mutations result in non-functional connexin hemichannels and gap junctions in C6 glioma cells. J. Cell Sci. 2006, 119 Pt 3, 532–541. [Google Scholar] [CrossRef]
  145. Ponsaerts, R.; De Vuyst, E.; Retamal, M.; D’Hondt, C.; Vermeire, D.; Wang, N.; De Smedt, H.; Zimmermann, P.; Himpens, B.; Vereecke, J.; et al. Intramolecular loop/tail interactions are essential for connexin 43-hemichannel activity. FASEB J. 2010, 24, 4378–4395. [Google Scholar] [CrossRef]
  146. Pal, J.D.; Liu, X.; Mackay, D.; Shiels, A.; Berthoud, V.M.; Beyer, E.C.; Ebihara, L. Connexin46 mutations linked to congenital cataract show loss of gap junction channel function. Am. J. Physiol. Cell Physiol. 2000, 279, C596–C602. [Google Scholar] [CrossRef] [PubMed]
  147. Ren, Q.; Riquelme, M.A.; Xu, J.; Yan, X.; Nicholson, B.J.; Gu, S.; Jiang, J.X. Cataract-causing mutation of human connexin 46 impairs gap junction, but increases hemichannel function and cell death. PLoS ONE 2013, 8, e74732. [Google Scholar] [CrossRef] [PubMed]
  148. Bergoffen, J.; Scherer, S.S.; Wang, S.; Scott, M.O.; Bone, L.J.; Paul, D.L.; Chen, K.; Lensch, M.W.; Chance, P.F.; Fischbeck, K.H. Connexin mutations in X-linked Charcot-Marie-Tooth disease. Science 1993, 262, 2039–2042. [Google Scholar] [CrossRef]
  149. Zheng, L.; Chenavas, S.; Kieken, F.; Trease, A.; Brownell, S.; Anbanandam, A.; Sorgen, P.L.; Spagnol, G. Calmodulin Directly Interacts with the Cx43 Carboxyl-Terminus and Cytoplasmic Loop Containing Three ODDD-Linked Mutants (M147T, R148Q, and T154A) that Retain alpha-Helical Structure, but Exhibit Loss-of-Function and Cellular Trafficking Defects. Biomolecules 2020, 10, 1452. [Google Scholar] [CrossRef]
  150. Jiang, J.X. Gap junctions or hemichannel-dependent and independent roles of connexins in cataractogenesis and lens development. Curr. Mol. Med. 2010, 10, 851–863. [Google Scholar] [CrossRef]
  151. Villanelo, F.; Carrasco, J.; Jensen-Flores, J.; Garate, J.A.; Perez-Acle, T. Simulations on Simple Models of Connexin Hemichannels Indicate That Ca2+ Blocking Is Not a Pure Electrostatic Effect. Membranes 2021, 11, 372. [Google Scholar] [CrossRef] [PubMed]
  152. Jagielnicki, M.; Kucharska, I.; Bennett, B.C.; Harris, A.L.; Yeager, M. Connexin Gap Junction Channels and Hemichannels: Insights from High-Resolution Structures. Biology 2024, 13, 298. [Google Scholar] [CrossRef] [PubMed]
  153. Zonta, F.; Buratto, D.; Crispino, G.; Carrer, A.; Bruno, F.; Yang, G.; Mammano, F.; Pantano, S. Cues to Opening Mechanisms From in Silico Electric Field Excitation of Cx26 Hemichannel and in Vitro Mutagenesis Studies in HeLa Transfectans. Front. Mol. Neurosci. 2018, 11, 170. [Google Scholar] [CrossRef] [PubMed]
  154. Lee, H.J.; Jeong, H.; Hyun, J.; Ryu, B.; Park, K.; Lim, H.H.; Yoo, J.; Woo, J.S. Cryo-EM structure of human Cx31.3/GJC3 connexin hemichannel. Sci. Adv. 2020, 6, eaba4996. [Google Scholar] [CrossRef] [PubMed]
  155. Tao, S.; Hulpiau, P.; Wagner, L.E.; Witschas, K.; Yule, D.I.; Bultynck, G.; Leybaert, L. IP3RPEP6, a novel peptide inhibitor of IP(3) receptor channels that does not affect connexin-43 hemichannels. Acta Physiol. 2024, 240, e14086. [Google Scholar] [CrossRef] [PubMed]
  156. Peracchia, C.; Wang, X.G.; Peracchia, L.L. Chemical gating of gap junction channels. Methods 2000, 20, 188–195. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, X.G.; Peracchia, C. Molecular dissection of a basic COOH-terminal domain of Cx32 that inhibits gap junction gating sensitivity. Am. J. Physiol. 1998, 275 Pt 1, C1384–C1390. [Google Scholar] [CrossRef]
  158. Zhou, Y.; Yang, W.; Lurtz, M.M.; Chen, Y.; Jiang, J.; Huang, Y.; Louis, C.F.; Yang, J.J. Calmodulin mediates the Ca2+-dependent regulation of Cx44 gap junctions. Biophys. J. 2009, 96, 2832–2848. [Google Scholar] [CrossRef]
  159. Peracchia, C.; Wang, X.G.; Peracchia, L.L. Behavior of chemical- and slow voltage-sensitive gating of connexin channels: The “cork” gating hypothesis. In Gap Junctions. Molecular Basis of Cell Communication in Health and Disease; Peracchia, C., Ed.; Academic Press: San Diego, CA, USA, 2000; pp. 271–295. [Google Scholar]
  160. Lissoni, A.; Wang, N.; Nezlobinskii, T.; De Smet, M.; Panfilov, A.V.; Vandersickel, N.; Leybaert, L.; Witschas, K. Gap19, a Cx43 Hemichannel Inhibitor, Acts as a Gating Modifier That Decreases Main State Opening While Increasing Substate Gating. Int. J. Mol. Sci. 2020, 21, 7340. [Google Scholar] [CrossRef]
  161. Ponsaerts, R.; Wang, N.; Himpens, B.; Leybaert, L.; Bultynck, G. The contractile system as a negative regulator of the connexin 43 hemichannel. Biol. Cell 2012, 104, 367–377. [Google Scholar] [CrossRef]
  162. Peracchia, C.; Peracchia, L.M.L. Calmodulin-Connexin Partnership in Gap Junction Channel Regulation-Calmodulin-Cork Gating Model. Int. J. Mol. Sci. 2021, 22, 13055. [Google Scholar] [CrossRef] [PubMed]
  163. Tran, O.; Kerruth, S.; Coates, C.; Kaur, H.; Peracchia, C.; Carter, T.; Torok, K. Ca2+-Dependent and -Independent Calmodulin Binding to the Cytoplasmic Loop of Gap Junction Connexins. Int. J. Mol. Sci. 2023, 24, 4153. [Google Scholar] [CrossRef] [PubMed]
  164. Torok, K.; Stauffer, K.; Evans, W.H. Connexin 32 of gap junctions contains two cytoplasmic calmodulin-binding domains. Biochem. J. 1997, 326 Pt 2, 479–483. [Google Scholar] [CrossRef] [PubMed]
  165. Ahmad, S.; Martin, P.E.; Evans, W.H. Assembly of gap junction channels: Mechanism, effects of calmodulin antagonists and identification of connexin oligomerization determinants. Eur. J. Biochem. 2001, 268, 4544–4552. [Google Scholar] [CrossRef] [PubMed]
  166. Mones, S.; Bordignon, B.; Peiretti, F.; Landrier, J.F.; Gess, B.; Bourguignon, J.J.; Bihel, F.; Fontes, M. CamKII inhibitors reduce mitotic instability, connexon anomalies and progression of the in vivo behavioral phenotype in transgenic animals expressing a mutated Gjb1 gene. Front. Neurosci. 2014, 8, 151. [Google Scholar] [CrossRef] [PubMed]
  167. Bihel, F.; Gess, B.; Fontes, M. CMTX Disorder and CamKinase. Front. Cell. Neurosci. 2016, 10, 49. [Google Scholar] [CrossRef] [PubMed]
  168. Caufriez, A.; Bock, D.; Martin, C.; Ballet, S.; Vinken, M. Peptide-based targeting of connexins and pannexins for therapeutic purposes. Expert. Opin. Drug Discov. 2020, 15, 1213–1222. [Google Scholar] [CrossRef] [PubMed]
  169. Buratto, D.; Donati, V.; Zonta, F.; Mammano, F. Harnessing the therapeutic potential of antibodies targeting connexin hemichannels. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166047. [Google Scholar] [CrossRef]
  170. AlFindee, M.N.; Subedi, Y.P.; Fiori, M.C.; Krishnan, S.; Kjellgren, A.; Altenberg, G.A.; Chang, C.T. Inhibition of Connexin Hemichannels by New Amphiphilic Aminoglycosides without Antibiotic Activity. ACS Med. Chem. Lett. 2018, 9, 697–701. [Google Scholar] [CrossRef]
Figure 1. ATP signalling mediated by connexin HCs. Typically, a local submicromolar increase in the cytosolic Ca2+ concentration ([Ca2+]c) triggers ATP release by connexin HCs of the same cell. The ATP molecules diffuse extracellularly and can activate ATP-gated P2X ionotropic receptors and G-protein-coupled P2Y receptors. The latter trigger a canonical transduction cascade mediated by the second messenger inositol 1,4,5-trisphosphate (InsP3). InsP3 binds to and opens its receptor (InsP3R) that in turn releases Ca2+ from the endoplasmic reticulum (ER) or the Golgi apparatus. The ensuing increase in [Ca2+]c promotes further HC openings, resulting in the propagation of a Ca2+ wave sustained by an ATP-induced ATP release mechanism. InsP3 can also diffuse through GJs, supporting the Ca2+ wave propagation.
Figure 1. ATP signalling mediated by connexin HCs. Typically, a local submicromolar increase in the cytosolic Ca2+ concentration ([Ca2+]c) triggers ATP release by connexin HCs of the same cell. The ATP molecules diffuse extracellularly and can activate ATP-gated P2X ionotropic receptors and G-protein-coupled P2Y receptors. The latter trigger a canonical transduction cascade mediated by the second messenger inositol 1,4,5-trisphosphate (InsP3). InsP3 binds to and opens its receptor (InsP3R) that in turn releases Ca2+ from the endoplasmic reticulum (ER) or the Golgi apparatus. The ensuing increase in [Ca2+]c promotes further HC openings, resulting in the propagation of a Ca2+ wave sustained by an ATP-induced ATP release mechanism. InsP3 can also diffuse through GJs, supporting the Ca2+ wave propagation.
Ijms 25 06594 g001
Figure 2. Ca2+ regulation of connexin HCs. Extracellular and cytosolic Ca2+ ions, in combination with the transmembrane voltage (Vm), finely regulate the molecular fluxes across the connexon pore. In healthy cells, this synergistic action allows the release of extracellular bursts of messenger molecules while preventing cytotoxic ionic leakage. An extracellular ring of Ca2+ ions stabilizes the HC closed conformation, while CaM and the actomyosin contractile system modulate the HC opening and closure mechanisms, respectively.
Figure 2. Ca2+ regulation of connexin HCs. Extracellular and cytosolic Ca2+ ions, in combination with the transmembrane voltage (Vm), finely regulate the molecular fluxes across the connexon pore. In healthy cells, this synergistic action allows the release of extracellular bursts of messenger molecules while preventing cytotoxic ionic leakage. An extracellular ring of Ca2+ ions stabilizes the HC closed conformation, while CaM and the actomyosin contractile system modulate the HC opening and closure mechanisms, respectively.
Ijms 25 06594 g002
Table 1. HC activity regulation by [Ca2+]e and [Ca2+]c. When available, the half maximal effective concentration (EC50) was reported. [Ca2+]e-free solution refers to an extracellular solution prepared without Ca2+ or with Ca2+ buffered by EGTA.
Table 1. HC activity regulation by [Ca2+]e and [Ca2+]c. When available, the half maximal effective concentration (EC50) was reported. [Ca2+]e-free solution refers to an extracellular solution prepared without Ca2+ or with Ca2+ buffered by EGTA.
Cx IsoformHC Regulation by [Ca2+]eHC Regulation by [Ca2+]c
Cx26Current ranging from a maximum at 0.01 mM to a minimum value at 10 mM [Ca2+]e, with an EC50 around 0.25 mM [78].ATP release increases with [Ca2+]c around 500 nM [8].
Cx30Current increases in [Ca2+]e-free solution [60].ATP release and dye uptake increases with [Ca2+]c around 500 nM [8].
Cx30.2/31.3ATP release increases in [Ca2+]e-free solution [12].Information not available.
Cx32Current ranging from a maximum at 0.5 mM to a minimum value at 5 mM [Ca2+]e, with an EC50 around 1.3 mM [72].Bell-shaped dependence of ATP release on [Ca2+]c, peaking at 500 nM [Ca2+]c [13].
Cx37Current ranging from a maximum at 0.02 mM to a minimum value at 1 mM [Ca2+]e, with an EC50 around 0.1 mM [79].Information not available.
Cx39Dye uptake increases in [Ca2+]e-free solution [80].Information not available.
Cx40HC pore size increases at [Ca2+]e < 10 μM [81].Information not available.
Cx43Dye uptake ranging from a maximum at 0.01 mM [Ca2+]e to a minimum value at 1 mM [Ca2+]e [63].Bell-shaped dependence of ATP release on [Ca2+]c, peaking at 500 nM [Ca2+]c [75].
Cx45Bi-sigmoidal dependence of the current on [Ca2+]e, with EC50 around 1 μM [Ca2+]e [82].Current increases with [Ca2+]c around 100 nM [82].
Cx46Current ranging from a maximum at 0.01 mM to a minimum value at 1 mM [Ca2+]e [71].Dye uptake increases upon [Ca2+]c elevation [85].
Cx50Dye leakage increases in [Ca2+]e-free solution [84].Information not available.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bayraktar, E.; Lopez-Pigozzi, D.; Bortolozzi, M. Calcium Regulation of Connexin Hemichannels. Int. J. Mol. Sci. 2024, 25, 6594. https://doi.org/10.3390/ijms25126594

AMA Style

Bayraktar E, Lopez-Pigozzi D, Bortolozzi M. Calcium Regulation of Connexin Hemichannels. International Journal of Molecular Sciences. 2024; 25(12):6594. https://doi.org/10.3390/ijms25126594

Chicago/Turabian Style

Bayraktar, Erva, Diego Lopez-Pigozzi, and Mario Bortolozzi. 2024. "Calcium Regulation of Connexin Hemichannels" International Journal of Molecular Sciences 25, no. 12: 6594. https://doi.org/10.3390/ijms25126594

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