Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2010 May 5;107(21):9897–9902. doi: 10.1073/pnas.1004552107

pH-dependent modulation of voltage gating in connexin45 homotypic and connexin45/connexin43 heterotypic gap junctions

Nicolas Palacios-Prado a, Stephen W Briggs a, Vytenis A Skeberdis a, Mindaugas Pranevicius b, Michael V L Bennett a,1, Feliksas F Bukauskas a,1
PMCID: PMC2906847  PMID: 20445098

Abstract

Intracellular pH (pHi) can change during physiological and pathological conditions causing significant changes of electrical and metabolic cell–cell communication through gap junction (GJ) channels. In HeLa cells expressing wild-type connexin45 (Cx45) as well as Cx45 and Cx43 tagged with EGFP, we examined how pHi affects junctional conductance (gj) and gj dependence on transjunctional voltage (Vj). To characterize Vj gating, we fit the gj–Vj relation using a stochastic four-state model containing one Vj-sensitive gate in each apposed hemichannel (aHC); aHC open probability was a Boltzmann function of the fraction of Vj across it. Using the model, we estimated gating parameters characterizing sensitivity to Vj and number of functional channels. In homotypic Cx45 and heterotypic Cx45/Cx43-EGFP GJs, pHi changes from 7.2 to ~8.0 shifted gj–Vj dependence of Cx45 aHCs along the Vj axis resulting in increased probability of GJ channels being in the fully open state without change in the slope of gj dependence on Vj. In contrast, acidification shifted gj–Vj dependence in the opposite direction, reducing open probability; acidification also reduced the number of functional channels. Correlation between the number of channels in Cx45-EGFP GJs and maximal gj achieved under alkaline conditions showed that only ~4% of channels were functional. The acid dissociation constant (pKa) of gj–pHi dependence of Cx45/Cx45 GJs was ~7. The pKa of heterotypic Cx45/Cx43-EGFP GJs was lower, ~6.7, between the pKas of Cx45 and Cx43-EGFP (~6.5) homotypic GJs. In summary, pHi significantly modulates junctional conductance of Cx45 by affecting both Vj gating and number of functional channels.

Keywords: cell-cell coupling, pH-dependent gating, EGFP, hemichannel, connexon


Changes in intracellular pH (pHi) take place under different physiological and pathological conditions, and H+ ions have a broad effect on cell function including cell–cell electrical and metabolic communication mediated by gap junctions (GJs) and paracrine signaling through nonjunctional/unapposed hemichannels (14). Modest pHi changes have been observed under normal physiological conditions [e.g., changes of neuronal activity or the resting potential (5, 6)], and greater changes occur under pathological conditions such as hypoxia, ischemia, or epilepsy (4, 7, 8).

During the last decade, significant progress has been made toward understanding the molecular mechanisms of pH-dependent modulation of GJs and hemichannels (2, 912). Several domains in the cytoplasmic loop and C terminus of connexin43 (Cx43) appear to be involved in pH-dependent gating (2, 9, 11, 12). Furthermore, pH-dependent interaction of connexins with other cytoplasmic proteins may be important in the remodeling of connexins and in protection from lesion spread after local ischemic injury (13, 14).

GJs provide channels with an inner diameter of ~1.4 nm between the interiors of the coupled cells. This link allows the spread of electrical potential and small metabolites. Each GJ channel is composed of two apposed hemichannels (aHCs) or connexons, each with six connexin subunits. aHCs can be homomeric (all subunits of the same connexin isoform) or heteromeric (containing more than one connexin isoform). GJ channels can be homotypic or heterotypic, i.e., formed by the same or different aHCs. Different connexin compositions can lead to differences in single-channel conductance, permselectivity, and/or asymmetric voltage gating (15, 16).

Each aHC has two mechanisms of gating in response to transjunctional voltage (Vj), the fast gate and the slow (or loop) gate (17) (Figs. S1 and S2). In addition, GJ channels can be gated by intracellular H+, Ca2+, posttranslational modifications, and a variety of chemical agents (18). The mechanisms by which these factors exert their effect on GJs remains unclear, although H+ and Ca2+ ions may both act through the slow Vj gate (11).

Here, we characterize pHi-dependent modulation of Cx45 GJs in HeLa cells expressing wild-type Cx45 or Cx45 tagged on its C terminus with green fluorescent protein (Cx45-EGFP). We also study heterotypic Cx45/Cx43-EGFP junctions to enable more accurate analysis of Cx45 aHCs. We found that much of the pHi-dependent change in junctional conductance (gj) could be explained by modulation of the voltage-gating properties. In addition, the number of functional channels (NF), i.e., those that can be gated by Vj, increased modestly during alkalinization to approach a maximum but decreased markedly during acidification. From estimates of the total number of GJ channels, gj as a function of pHi, and single-channel conductance, we calculate that at alkaline pHi only ~4% of the GJ channels in junctional plaques (JPs) are functional; the remaining 96% appear to be permanently closed (Fig. S2). Thus, pHi affects gj by shifting the gj–Vj relation of the aHCs along the Vj axis as well as by altering the number of channels that can open.

Results

pHi-Dependent Modulation of the Conductance of Cx45 GJs.

Junctional conductance (gj) between HeLa cells expressing Cx45 or Cx45-EGFP was measured by dual voltage-clamp and application of repeated Vj ramps, 600 ms in duration, from +10 to −10 mV (Fig. 1A, Inset). In the Cx45 cell pairs studied, gj under control conditions (pHi = 7.2) varied between 6 and 47 nS (n = 29). In HeLaCx45-EGFP cell pairs, gj under control conditions varied between 1.4 and 76 nS (n = 18), and, in general, gj was higher in cell pairs with larger JPs visualized by their EGFP fluorescence.

Fig. 1.

Fig. 1.

pHi-dependent modulation of gj in homotypic Cx45 GJs. (A) Junctional conductance (gj) measurements in a HeLaCx45 cell pair at different pHis modulated by exposure to CO2 and NH4Cl. Repeated Vj ramps of ±10 mV in amplitude and 600 ms in duration (Inset) were used to measure gj. (B) Normalized gj–pHi relations of GJs formed of Cx45 (circles) and of Cx45-EGFP (squares) with BCECF in the pipette to determine pHi; these relations are superimposable. Triangles show gj–pHi dependence for experiments in which the pipette solution was buffered at different pHs and BCECF was omitted; pHo remained constant (7.4); gj was normalized to the maximum value at alkaline pHi. Data for Cx45 (circles) were fit by the Hill equation (solid black curve); pKa = 6.96 ± 0.03; Hill coefficient n = 1.43 (n = 12). Thinner dashed lines show 95% confidence interval (CI).

To examine gj as a function of pHi, we used CO2 or ammonium chloride (NH4Cl) to reduce or increase pHi, respectively. Cells were exposed to modified Krebs-Ringer (MKR) solutions bubbled with CO2 or with added NH4Cl, and the unesterified form of the ratiometric fluorescent probe, 2,7-bis(2-carboxyethyl)-5(6)-carboxyfluorescein (BCECF), was dialyzed into the cells from the recording pipettes to measure pHi (Fig. 1A) (19). Application of MKR solution bubbled with 5% CO2 decreased pHi to ~6.3, resulting in almost complete uncoupling. CO2 decreased extracellular pH (pHo) as well, but gj has been shown to be insensitive to external acidification (20). The small increase in gj (arrow in Fig. 1A; also see Fig. 4B) as pHi starts to decrease has been described (21). Application of 15 mM NH4Cl increased pHi from 7.2 ± 0.1 to 8.1 ± 0.2 (n = 18) and increased gj on average to 1.8 times its value at pHi = 7.2. Fitting the Hill equation to the gj–pHi relation for Cx45 (circles in Fig. 1B) yielded an acid dissociation constant (pKa) equal 6.96 ± 0.03 and Hill coefficient n = 1.43 (n = 12). Values for Cx45-EGFP (squares in Fig. 1B) were not significantly different (pKa = 6.90 ± 0.04 and Hill coefficient n = 1.45, n = 4). To test whether BCECF itself could affect our measurements, we used pipette solutions without BCECF and buffered to different pHs. After opening of the patches, gj was measured after reaching a steady state. Then the cells were exposed to MKR solution containing a high concentration of NH4Cl (20 mM) at pHo ~8, and the gj–pHi relation was normalized to the maximum gj. The observed gj–pHi dependence (triangles in Fig. 1B) overlapped the gj–pHi relation determined using the BCECF probe, a convergence validating both approaches.

Fig. 4.

Fig. 4.

pHi-dependent modulation of gj in heterotypic Cx45/Cx43-EGFP GJs. (A) Fluorescence image of a HeLa cell pair forming heterotypic Cx45/Cx43-EGFP GJs are enlarged in Inset. (B) pHi and gj measured in a Cx45/Cx43-EGFP cell pair exposed to CO2 at the indicated percentages and at increasing concentrations of NH4Cl alone or combined with alkaline MKR solution (pHo = 8.3) and finally with 3% CO2. (C) Junctional conductance (gj)–pHi dependence of Cx45/Cx43-EGFP GJs (white circles are from the experiment shown in B and gray circles are from other five experiments). The red line shows fit of the Hill equation to the data; pKa = 6.70 ± 0.03, Hill coefficient n = 1.6 (n = 6). Dashed lines show 95% CI. (D) Po–pHi dependence of Cx45 (green) and Cx43-EGFP (blue) homotypic GJs calculated from data shown Fig. 1 and Fig. S4, respectively. Dashed green and blue lines are square roots of Po–pHi dependence of Cx45 and Cx43-EGFP homotypic GJs, respectively. (E) Po–pHi dependence of Cx45 (green), Cx43-EGFP (blue), and Cx45/Cx43-EGFP (red) GJs. The black line shows the product of Po,Cx45–pHi and Po,Cx43–pHi plots shown in D (predicted). (F) Junctional conductance (gj)–Vj plots (black lines) calculated by applying pairs of Vj ramps (35 s duration) from 0 to +100 and to −100 mV at the numbered times shown on the gj trace in B. Gray lines show calculated gj–Vj plots fit by the S4SM to the experimental data. (G) Probabilities of channels being in the O-O state at times of ramps #1, #3, and #6 obtained from fits to gj–Vj plots with indicated pHis and numbers corresponding to those shown in F. (H) Relation between gj normalized to gj at Vj = 0 under control conditions (gj0, norm.) and Vo, Cx45. (I) Relation between NF normalized to NF at pH = 7.2 and Vo,Cx45. Acidic conditions are shown in light pink, and alkaline conditions are shown in light blue relative to pHi = 7.2. Solid lines in H and I are fits of a sigmoidal equation, y = a/(1+e−(x−xo)/b). Dashed lines show 95% CI.

pH-Dependent Modulation of Vj Gating in Cx45 GJ Channels.

We examined gj–Vj dependence of homotypic Cx45 GJs during acidification (Fig. 2) and alkalinization (Fig. 3). Because homotypic junctions typically demonstrate symmetric gj–Vj dependence, we measured junctional current (Ij) only in response to 31-s Vj ramps of negative polarity, from 0 to −110 mV (Figs. 2A and 3A). The experimental gj–Vj relations calculated from Ij and Vj records are hemibell shaped with gj maximal at Vj = 0 and markedly lower at more negative Vjs (Figs. 2B and 3B, black curves). During acidification, gj was reduced (Fig. 2 A and B); during alkalinization, gj was increased (Fig. 3 A and B). To analyze Vj-gating properties, we fitted experimental gj–Vj plots with a stochastic four-state model (S4SM) (Fig. S1) (22). The S4SM assumes that the voltage across each aHC in the junction depends on the conductance of the aHC in series (“contingent gating”) (23) and allows us to define parameters characterizing Vj gating according to a Boltzmann relation, namely Vo,H (voltage across an aHC at which its open probability, Po,H, is 0.5) and AH (coefficient characterizing the steepness of Po,H changes as a function of VH). In addition, the S4SM allows us to define the actual number of Cx45 GJ channels with unitary conductance (γo) of 32 pS and residual conductance of ~4 pS (24) that should be functional at any given time to explain the observed gj (Fig. S2); NF is the number of GJ channels with both slow gates open and with their fast Vj-gates open or closed, depending on pHi and Vj. Thus, during the fitting process, Vo,H, AH, and NF were free parameters. The residual conductance and its rectification factor were additional free parameters.

Fig. 2.

Fig. 2.

Vj gating during acidification of homotypic Cx45 GJs. (A) Ij measured in a HeLa Cx45 cell pair in response to repeated Vj ramps (31 s duration) from 0 to −100 mV before and during exposure to 5% CO2. Vj steps of −10 mV were used to measure gj between Vj ramps. (B) Experimental gj–Vj plots (in black) were fit by the S4SM (in gray) assuming that Vj gating was symmetric around Vj = 0. (C) Relation between gj at Vj = 0 (gj0; normalized to control/initial conditions) and Vo,H. (black and white circles). (D) Relation between NF normalized to control/initial conditions and Vo,H (black and white circles). Solid lines in C and D are curves fit from combined data in Fig. 2 and Fig. 3, obtained using a sigmoidal equation, y = a/(1+e−(x−xo)/b) (n = 12). Dashed lines show 95% CI. White circles are from the experiment shown in A. (Ea–c) Probabilities of channels to dwell in O-O, C-O, O-C, and C-C states (PS) depending on Vj and calculated for the three Ij records obtained during Vj ramps a–c marked with gray rectangles in A; calculated values of Vo,H are indicated.

Fig. 3.

Fig. 3.

Vj gating during alkalinization of homotypic Cx45 GJs. (A) Ij measured in a HeLaCx45 cell pair in response to repeated Vj ramps (31 s duration) from 0 to −100 mV before and during exposure to NH4Cl alone or to NH4Cl combined with alkaline conditions (pHo = 8 or 8.3). Vj steps of −10 mV were used to measure gj between Vj ramps. (B) Experimental gj–Vj plots (in black) obtained at Vj ramps marked with gray rectangles in A were fitted to the S4SM (in gray) assuming that Vj gating was symmetric around Vj = 0. (C) Relation between gj at Vj = 0 (gj0; normalized to control/initial conditions) and Vo,H (black and white circles). (D) Relation between NF normalized to control/initial conditions and Vo,H (black and white circles). Solid lines in C and D are fits to combined data from Fig. 2 and Fig. 3, obtained using a sigmoidal equation, y = a/(1+e−(x−xo)/b) (n = 12). Dashed lines show 95% CI. White circles are from the experiment shown in A. (E) Probabilities of channels to dwell in O-O, C-O, O-C, and C-C states (PS) depending on Vj and calculated for the three Ij records obtained at Vj ramps ac marked with gray rectangles in A; calculated values of Vo,H are indicated on each graph.

Fitting of experimental gj–Vj plots (gray lines in Figs. 2B and 3B) was performed assuming that each experimental gj–Vj relation (black lines in Figs. 2B and 3B) is symmetric around Vj = 0 (Table S1 shows the parameters obtained to fit curves in Figs. 2B and 3B). In the S4SM we assumed the single aHC conductance (γo,H) for Cx45 was 64 pS (24) and that the residual conductance (γres) rectified so as to increase exponentially when the cytoplasmic side of the Cx45 aHC was more negative (legend in Fig. S1) (15).

The uncoupling effect of acidification with 5% CO2 decreased Vo,H by ~35 mV and decreased gj and NF to near zero, whereas AH remained virtually constant (Fig. 2C and Table S1). In contrast, alkalinization increased gj 1.8-fold (Fig. 3B) and increased Vo,H by ~40 mV (Fig. 3C); there was a small increase in NF, and AH remained virtually constant (Table S1). The same effects on Vj gating were found in all seven experiments using intracellular acidification and all six experiments using alkalinization, including the one shown in Fig. S3, where both CO2 and NH4Cl were applied in the same experiment.

Vj dependence of the probabilities (PS) of the four possible channel states [open-open (O-O), both aHCs open; open-closed (O-C) and closed-open (C-O), one aHC open and one closed; and closed-closed (C-C), both aHCs closed (Fig. S2)] calculated with S4SM at acid, alkaline, and control pHi is shown in Figs. 2E a–c and 3E a–c. The decrease in gj during acidification was the result of a reduction of both PO-O (Figs. 2E a–c), caused by decrease in Vo,H (Fig. 2C) and NF (Fig. 2D). In contrast, increase in gj during alkalinization was mainly the result of an increase of PO-O to ~1 (Fig. 3Ec), caused by increase in Vo,H (Fig. 3C), and a slight increase in NF (Fig. 3D).

pHi-Dependent Modulation of gj and Vj Gating in Cx45/Cx43-EGFP Heterotypic GJ Channels.

We examined the gj–pHi dependence of heterotypic Cx45/Cx43-EGFP GJs to study pHi sensitivity of Cx45 aHCs and to test the hypothesis that pHi sensitivity of aHCs is similar in homotypic and heterotypic GJs. Cx45 GJs are more sensitive to pHi and Vj than Cx43-EGFP GJs, and the Po,H of the less sensitive Cx43-EGFP aHC changed relatively little over the pHi and Vj range that changed the Po,Cx45 aHC from high to low. For this reason we could evaluate the pHi and Vj dependence of Cx45 aHCs at higher resolution in heterotypic than in homotypic junctions.

Homotypic Cx43 GJs show a pKa between 6.6 and 6.7 (9, 25). We found that the pKa of Cx43-EGFP GJs was slightly more acidic (pKa = 6.50 ± 0.03, Hill coefficient n = 1.98, n = 6) (Fig. S4). Thus, the pKa of Cx43-EGFP is ~0.5 units more acidic than that of Cx45 (pKa≈7) (Fig. 1). Cell pairs forming heterotypic Cx45/Cx43-EGFP GJs were selected based on EGFP expression in only one cell and the presence of one or more JPs between the cells (Fig. 4A). Junctional conductance (gj) was measured by repeated application of Vj ramps from +10 to −10 mV and 600 ms in duration. Under control conditions (pHi = 7.2), gj varied between 15 and 71 nS (n = 18). To study pHi effects on gj, we perfused cells with MKR solution bubbled with different percentages of CO2 or containing different concentrations of NH4Cl (Fig. 4B). NH4Cl increased gj, which reached a plateau when exposed to MKR solutions with pHo = 8.3 and containing 15 mM of NH4Cl that increased pHi to 8.3 ± 0.1 (n = 6). Here and below, we used alkaline MKR to increase pHi; at neutral pHo we would have needed 30 mM or more NH4Cl to reach pHi= 8.3, which might have been an excessive increase in osmolarity. The best fit of gj–pHi dependence by the Hill equation was obtained with pKa = 6.70 ± 0.03 and Hill coefficient n = 1.6 (n = 6) (Fig. 4C). Hypothetically, the gj–pHi dependence of GJs results from the pHi sensitivities of the two aHCs in series, and Po = Po,H1·Po,H2 (20). We estimated the Po,H–pHi dependence for Cx45 and Cx43-EGFP aHCs (Fig. 4D, dashed lines) by calculating the square root of Po at each pHi for homotypic GJs shown in Fig. 1B and Fig. S4, respectively. The Po–pHi relation and pKa for heterotypic Cx45/C43-EGFP GJs falls between the Po–pHi relations and pKas of the homotypic Cx45 and Cx43-EGFP GJs (pKa,Cx45 ≈7; pKa,Cx45/Cx43-EGFP ≈ 6.7; pKa,Cx43-EGFP ≈ 6.5). Thus, to a first approximation, the product of Po,H–pHi dependence of Cx45 and Cx43-EGFP aHCs predicts the Po–pHi dependence of heterotypic GJs (Fig. 4E), sug-gesting that the series aHCs in heterotypic GJs respond independently to pHi.

To examine pHi effects on voltage-gating properties of Cx45/Cx43-EGFP GJs, we determined gj–Vj relations at different pHi by applying slow Vj ramps from 0 to +100 and −100 mV at the times indicated by the numbers on the gj trace in Fig. 4B. In accordance with earlier reports (24, 26), the gj–Vj dependence at pHi = 7.2 was highly asymmetric around Vj = 0 (Figs. 4F and 5B). Both Cx45 and Cx43-EGFP aHCs exhibit negative gating polarity, i.e., they close at relative negativity on their cytoplasmic side (24). Vjs for which the Cx45 side is relatively negative (which we define as positive Vj for Cx45/Cx43 heterotypic junctions) tend to close the Cx45 aHCs and open the Cx43 aHCs (Figs. 4F and 5B; gray lines are curves obtained with parameters shown in Table S2) (17). Vjs of the opposite sign tend to close the Cx43-EGFP aHCs and open the Cx45 aHCs. In addition, Cx43-EGFP lacks the fast-gating mechanism and is significantly less Vj sensitive than Cx45 (27). Finally, in heterotypic Cx45/Cx43-EGFP channels with both aHCs open, a larger fraction of an applied Vj acts on the Cx45 aHC, because its conductance is ~one-fourth that of the Cx43 aHC (24), and consequently Cx45 aHCs are more Vj sensitive in Cx45/Cx43-EGFP GJs than in Cx45/Cx45 GJs (22). As a result of these factors, Vj gating in Cx45/Cx43-EGFP GJs is determined mainly by the Cx45 aHCs at positive Vjs, and the rightward shift of Vj sensitivity without change in slope along the Vj axis of heterotypic GJs during alkalinization (evident from data shown in Fig. 4F) is ascribable almost entirely to modulation of the Cx45 aHCs. Hence, Cx45 aHCs in homotypic GJs and heterotypic GJs with Cx43-EGFP exhibit similar modulation of Vj gating by pHi. Figs. 4G and 5C show probabilities of the channels being in the O-O state during alkalinization and acidification, respectively, at times corresponding to those indicated in Figs. 4F and 5B. PO-O–Vj plots show that alkalinization enhances PO-O over the entire Vj range, accounting for the rightward shift of the gj–Vj relation (Fig. 4F), whereas acidification reduces PO-O and NF, and shifts the PO-O–Vj relation to the left, explaining the decrease in PO-O and gj. At acid pHi and Vj < −30 mV, PO-O decreases, presumably because of the Cx43 aHC closing. Figs. 4H and 5D show relations between gj0 (normalized to gj at Vj = 0 and pHi ≈ 7.2) and Vo,Cx45. Figs. 4I and 5E show relations between NF (normalized to its value at pHi ≈ 7.2) and Vo,Cx45.. These relations were obtained from fits of gj–Vj plots shown in Figs. 4F and 5B, respectively, and support the conclusion reached in Cx45 homotypic GJs that acidification decreases gj because of the decrease in Vo,H and NF.

Fig. 5.

Fig. 5.

Vj gating during acidification of heterotypic Cx45/Cx43-EGFP GJs. (A) Ij measured in a Cx45/Cx43-EGFP cell pair in response to repeated Vj ramps (25 s duration) from 0 to −100 mV and to +100 mV, and short (500-ms) steps of 20 mV during application of 5% CO2. (B) Junctional conductance (gj)–Vj data (black lines), calculated from Vj and Ij records marked with numbers from 1 to 8 shown in A, were fit by the S4SM (in gray). (C) Probability that GJs dwell in the O-O state as a function of Vj; attached numbers correspond to those shown in A and B. (D) Relation between gj normalized to its value under control conditions at Vj = 0 (gj0, norm.) and Vo,Cx45; numbers correspond to those shown in A and B. (E) Relation between NF normalized to NF at pH = 7.2 and Vo,Cx45. Solid lines in D and E are fit to the data using a linear regression of the second order. Dashed lines show 95% CI.

Functional Efficiency of Cx45-EGFP GJ Channels.

We have reported that only ~1/10 of Cx43-EGFP and ~1/100 of Cx57 channels assembled into JPs are open at any one time at Vj = 0, and we called this fraction (K) “functional efficiency” (19, 28). To estimate K for Cx45-EGFP GJs, we used the same approach. In brief, we selected cell pairs with JPs that were oriented parallel to the focal plane and could be viewed en face (Fig. 6A). To obtain fluorescence per unit area without edge effects, we selected JPs that were sufficiently large so that the fluorescence was uniform in a central region. We assessed fluorescence per unit area (FJP) in this region in arbitrary fluorescent units (a.u.) and subtracted background fluorescence outside the JP. We assumed that each Cx45-EGFP channel occupied 100 nm2 (corresponding to 10-nm center-to-center spacing in a square array and 104 channels per μm2) as an approximation of values seen in atomic force microscopy (29, 30). Thus, we assessed the fluorescence produced by a single GJ channel (Fγ) from the ratio FJP/10,000; on average Fγ = 0.173 ± 0.001 a.u. (n = 19).

Fig. 6.

Fig. 6.

Functional efficiency of HeLaCx45-EGFP GJs. (A) Fluorescence image of a region where two cells overlap and form a JP oriented parallel to the focal plane. The area encircled by the dashed line is the region of interest (ROI) in which fluorescence intensity could be measured without edge effects because of the large size of the JP. (B) Image of a HeLaCx45 cell pair forming a JP enclosed in an ROI. Solid lines indicate cell borders. (C) The relation between NFmax estimated from gj measurements and NT. Linear regression (solid line) shows that under alkaline conditions NFmax was linearly related to NT with slope of 0.039 ± 0.03. Thus, only ~4% of GJ channels are functional.

In combined imaging and electrophysiological studies, we measured gj and the total fluorescence intensity (FT) of JPs independent of their spatial orientation. FT was estimated by measuring the total fluorescence in the region of interest (ROI) enclosing a JP (dashed ellipse in Fig. 6B). To collect all light including that from regions that were not in the exact focus, the ROI was made several fold larger than the size of the JPs (28). The total number of GJ channels present in each JP (NT) was determined from the ratio, NT = FT/Fγ. JPs with ~4.6 × 103 to 75 × 103 GJ channels (n = 14) based on fluorescence ranged from ~1–4 μm in diameter. During exposure to MKR solution at pHo = 8.3 and containing 10 mM NH4Cl, PO-O at Vj = 0 was equal to 0.98 ± 0.03 (n = 6) (Fig. 3Ec and Fig. S3Ec), and NF approached a maximum, NFmax, which is equal to maximum gj (gjmax)/γο. We hypothesize that PO-O for these channels was ~1 and that PO-O for the remaining, nonfunctional channels was ~0. Thus, the fraction of functional channels is the ratio of NFmax to the total number of channels, K = NFmax/NT. We found that under alkaline conditions for GJs ranging from 102 to 3,160 channels, NFmax was linearly related to NT with K = 0.039 ± 0.003 (n = 14) (Fig. 6C), and the fraction of functional channels is ~1/25.

Discussion

Here we propose that an increase of gj between Cx45-expressing cells with alkalinization from normal pHi is caused mainly by an increase in open probability, PO-O, of the functional channels. Acidification-induced full uncoupling is approached by decrease in both PO-O and NF. The parameter AH characterizing steepness of gj dependence on VH, remained virtually constant during pHi changes, suggesting that the gating charge involved in voltage sensing did not change significantly. Generally, gj–Vj dependence has been used to compare and characterize GJ channels formed by different Cx isoforms (31). Here we demonstrated that Vj gating can be modulated by pHi (see also ref. 32). For example, Cx45 GJs, which are highly Vj sensitive at normal pHi, can be modulated by intracellular alkalinization, so that their gj–Vj dependence resembles that of connexin26, which is relatively Vj insensitive around Vj = 0. Similar changes were reported for connexin 57 (Cx57) (19). Fluorescence imaging showed that the size of Cx45-EGFP and Cx43-EGFP homotypic GJs remained constant during exposure to and recovery from CO2 or NH4Cl (for Cx43-EGFP, see ref. 27). Furthermore, the pHi-dependent changes in gj over the time intervals measured appeared too fast and reversible to result from de novo formation or internalization of GJ channels. Fitting analysis showed that with increasing alkalinization PO-O increases asymptotically to a maximum that we hypothesize is unity at Vj = 0. If so, only ~1/25 of Cx45-EGFP channels assembled into JPs are functional, i.e., can be activated by changes in Vj and pHi (Fig. 6). The presence of such a small fraction of functional channels is not unique to Cx45 but also applies to Cx43 (28) and Cx57 (19), and their functionality may be determined by factors such as time needed for maturation after docking of two hemichannels, interaction with other proteins, and phosphorylation, which may move channels between functional and nonfunctional populations. Moreover, channels not mediating cell–cell coupling may be involved in other functions.

We found that pKa of Cx45/Cx43-EGFP heterotypic GJs is equal to 6.7, which is between pKas measured for GJs formed by Cx45 (Fig. 1B; pKa ≈ 7.0) and Cx43-EGFP (Fig. S4B; pKa ≈ 6.5). pKa reported for wild-type Cx43 measured in Xenopus oocytes (9) and Novikoff hepatoma cells (21) is 6.7; thus, attachment of EGFP to Cx43 C terminus appears to acidify pKa,Cx43-EGFP by ~0.2 units as well as to eliminate the fast-gating mechanism (27). In the pHi range of 6.7–8.0, Cx45 aHCs largely determine gj of Cx45/Cx43-EGFP heterotypic GJs. Theoretically, if the two aHCs in series act independently, the gj–pHi dependence of GJs results from pHi sensitivities of both aHCs in series, and pKa should be ~6.7, as observed experimentally (Fig. 4E). Thus, in these heterotypic junctions, at least, docking of unapposed hemi-channels appears not to affect their sensitivity to pHi.

Similarly, Vj gating in Cx45/Cx43-EGFP GJs should be determined mainly by the Cx45 aHC for two reasons: (i) Cx45 is more Vj sensitive than Cx43-EGFP in homotypic junctions (24), and (ii) most of the Vj drops across the Cx45 aHC because of its ~3.6-fold lower conductance, resulting in an increase and decrease of Vj gating of Cx45 and Cx43-EGFP aHCs, respectively (22). In Cx45/Cx43-EGFP heterotypic GJs, sensitivity to Vj around Vj = 0 is caused mainly by changes in the Cx45 aHC.

Changes in pHi have important effects on cell function. For example, in cardiac muscle changes in pHi influence normal tissue function by affecting excitability, contractility, and intercellular communication through GJs (33). This pH-dependent modulation of GJ channels may be critical under severe pathological conditions, because the uncoupling effect of low pHi may reduce propagation of metabolic stress from damaged cells to healthy neighbors (34, 35) or prevent rescue of stressed cells by healthier neighbors, as discussed later. Brief periods of global ischemia can produce fast intracellular acidosis of ~0.4 pH units in isolated ferret hearts (36). Increase in impulse frequency rate from 1 to 3 Hz causes a decrease in pHi by ~0.3 units in isolated sheep Purkinje fibers (37). pHi decreases more dramatically during global ischemia in a perfused rat heart model (38) when pHi drops to ~6.2, which accounts for much of the observed failure of contraction (36). A pHi decrease from 7.2 to 6.8 in cells expressing Cx45 would reduce gj more than twofold. This change would affect cell–cell metabolic communication, electrical signaling, and, consequently, the conduction of excitation in tissues where Cx45 is expressed extensively, e.g., the conduction system of the heart (39), smooth muscle cells of blood vessels (40), and certain neurons (41). Furthermore, Cx45 knock-out mice die in utero because of atrioventricular conduction block, impairment of atrial contraction, and severe dilation of the heart (42, 43).

Ischemia can result in local acidosis where capillary perfusion is greatly reduced, and neighboring less acid-loaded regions can help maintain pH by diffusion of hydrogen ions through the extracellular and intracellular compartments. Connexin-based GJ channels and hemichannels may serve as conduits for intracellular acid dissipation, especially because the classical sarcolemmal routes for acid efflux appear to be inhibited (44). Studies of normal/ischemia border regions in isolated preparations of the heart showed that preservation of ischemic cells by neighboring cells exposed to normal perfusion is greater in the longitudinal than in transverse direction of the fibers, consistent with ~7- to 10-fold stronger cell–cell coupling in the longitudinal direction (45). However, under severe ischemic conditions GJ communication is blocked (46), and this blockage may limit both the spread of damage and the cell-rescuing/preservation effect of GJs. Thus, during mild ischemia GJs remain open to preserve ischemia-affected cells, but when ischemia worsens, GJs shut down, a response that may isolate damaged areas and improve organ survival.

In summary, modulation at alkaline pH of cell–cell coupling of Cx45 homotypic and Cx45/Cx43 heterotypic GJ channels can be explained largely by changes in Vo,H with little change in AH and NF. Acidification decreases the number of functional channels and shifts Vo,H so as to reduce PO-O, while AH remains constant. The decrease in Vo,H appears continuous over the entire pH range investigated. i.e., ~6–8, suggesting a single mechanism. Cx45 GJ channels contain both fast and slow gates. Although we cannot yet clearly separate the effects of fast and slow gates on Vj- and pHi-dependent gating, we hypothesize that pHi affects NF through modulation of the slow gate, whereas changes in Vj gating are largely determined by the fast gate.

Methods

SI Text includes details of the experimental procedures.

Experiments with homotypic junctions were performed on HeLa cells transfected with wild-type Cx45, Cx45-EGFP, or Cx43-EGFP. Experiments with heterotypic junctions were performed on cocultures of HeLa cells expressing Cx45 or Cx43-EGFP. Junctional conductance (gj) was measured using the dual whole-cell voltage clamp method (15). Fluorescence signals were acquired using UltraVIEW software for image acquisition and analysis (Perkin-Elmer Life Sciences). For ratiometric pHi measurement, the unesterified form of BCECF (10 μM) was introduced into the cells through patch pipettes in whole-cell voltage-clamp mode. Dye was excited alternately with low-intensity 436-nm and 500-nm light for 0.5 s every 15 s (to minimize photobleaching), and the emitted light was filtered at 540 nm.

Supplementary Material

Supporting Information

Acknowledgments

We thank Angele Bukauskiene for excellent technical assistance. This work was supported by National Institutes of Health Grants NS036706 and HL084464 to F.F.B. and NS045287 to M.V.L.B.

Footnotes

The authors declare no conflict of interest.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1004552107/-/DCSupplemental.

References

  • 1.Demaurex N. pH Homeostasis of cellular organelles. News Physiol Sci. 2002;17:1–5. doi: 10.1152/physiologyonline.2002.17.1.1. [DOI] [PubMed] [Google Scholar]
  • 2.Delmar M, Coombs W, Sorgen P, Duffy HS, Taffet SM. Structural bases for the chemical regulation of connexin43 channels. Cardiovasc Res. 2004;62:268–275. doi: 10.1016/j.cardiores.2003.12.030. [DOI] [PubMed] [Google Scholar]
  • 3.Sáez JC, Retamal MA, Basilio D, Bukauskas FF, Bennett MVL. Connexin-based gap junction hemichannels: Gating mechanisms. Biochim Biophys Acta. 2005;1711:215–224. doi: 10.1016/j.bbamem.2005.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Obara M, Szeliga M, Albrecht J. Regulation of pH in the mammalian central nervous system under normal and pathological conditions: Facts and hypotheses. Neurochem Int. 2008;52:905–919. doi: 10.1016/j.neuint.2007.10.015. [DOI] [PubMed] [Google Scholar]
  • 5.Chesler M, Kaila K. Modulation of pH by neuronal activity. Trends Neurosci. 1992;15:396–402. doi: 10.1016/0166-2236(92)90191-a. [DOI] [PubMed] [Google Scholar]
  • 6.Lyall V, Biber TU. Potential-induced changes in intracellular pH. Am J Physiol. 1994;266:685–696. doi: 10.1152/ajprenal.1994.266.5.F685. [DOI] [PubMed] [Google Scholar]
  • 7.Diarra A, Sheldon C, Brett CL, Baimbridge KG, Church J. Anoxia-evoked intracellular pH and Ca2+ concentration changes in cultured postnatal rat hippocampal neurons. Neuroscience. 1999;93:1003–1016. doi: 10.1016/s0306-4522(99)00230-4. [DOI] [PubMed] [Google Scholar]
  • 8.Madden JA, Keller PA, Kleinman JG. Changes in smooth muscle cell pH during hypoxic pulmonary vasoconstriction: A possible role for ion transporters. Physiol Res. 2000;49:561–566. [PubMed] [Google Scholar]
  • 9.Stergiopoulos K, et al. Hetero-domain interactions as a mechanism for the regulation of connexin channels. Circ Res. 1999;84:1144–1155. doi: 10.1161/01.res.84.10.1144. [DOI] [PubMed] [Google Scholar]
  • 10.Trexler EB, Bukauskas FF, Bennett MVL, Bargiello TA, Verselis VK. Rapid and direct effects of pH on connexins revealed by the connexin46 hemichannel preparation. J Gen Physiol. 1999;113:721–742. doi: 10.1085/jgp.113.5.721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Peracchia C. Chemical gating of gap junction channels; roles of calcium, pH and calmodulin. Biochim Biophys Acta. 2004;1662:61–80. doi: 10.1016/j.bbamem.2003.10.020. [DOI] [PubMed] [Google Scholar]
  • 12.Hirst-Jensen BJ, Sahoo P, Kieken F, Delmar M, Sorgen PL. Characterization of the pH-dependent interaction between the gap junction protein connexin43 carboxyl terminus and cytoplasmic loop domains. J Biol Chem. 2007;282:5801–5813. doi: 10.1074/jbc.M605233200. [DOI] [PubMed] [Google Scholar]
  • 13.Duffy HS, et al. Regulation of connexin43 protein complexes by intracellular acidification. Circ Res. 2004;94:215–222. doi: 10.1161/01.RES.0000113924.06926.11. [DOI] [PubMed] [Google Scholar]
  • 14.Barker RJ, Price RL, Gourdie RG. Increased association of ZO-1 with connexin43 during remodeling of cardiac gap junctions. Circ Res. 2002;90:317–324. doi: 10.1161/hh0302.104471. [DOI] [PubMed] [Google Scholar]
  • 15.Rackauskas M, et al. Gating properties of heterotypic gap junction channels formed of connexins 40, 43, and 45. Biophys J. 2007;92:1952–1965. doi: 10.1529/biophysj.106.099358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Palacios-Prado N, Bukauskas FF. Heterotypic gap junction channels as voltage-sensitive valves for intercellular signaling. Proc Natl Acad Sci USA. 2009;106:14855–14860. doi: 10.1073/pnas.0901923106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bukauskas FF, Verselis VK. Gap junction channel gating. Biochim Biophys Acta. 2004;1662:42–60. doi: 10.1016/j.bbamem.2004.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Harris AL. Emerging issues of connexin channels: Biophysics fills the gap. Q Rev Biophys. 2001;34:325–472. doi: 10.1017/s0033583501003705. [DOI] [PubMed] [Google Scholar]
  • 19.Palacios-Prado N, Sonntag S, Skeberdis VA, Willecke K, Bukauskas FF. Gating, permselectivity and pH-dependent modulation of channels formed by connexin57, a major connexin of horizontal cells in the mouse retina. J Physiol. 2009;587:3251–3269. doi: 10.1113/jphysiol.2009.171496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Spray DC, Harris AL, Bennett MVL. Gap junctional conductance is a simple and sensitive function of intracellular pH. Science. 1981;211:712–715. doi: 10.1126/science.6779379. [DOI] [PubMed] [Google Scholar]
  • 21.Lazrak A, Peracchia C. Gap junction gating sensitivity to physiological internal calcium regardless of pH in Novikoff hepatoma cells. Biophys J. 1993;65:2002–2012. doi: 10.1016/S0006-3495(93)81242-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Paulauskas N, Pranevicius M, Pranevicius H, Bukauskas FF. A stochastic four-state model of contingent gating of gap junction channels containing two “fast” gates sensitive to transjunctional voltage. Biophys J. 2009;96:3936–3948. doi: 10.1016/j.bpj.2009.01.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Harris AL, Spray DC, Bennett MVL. Kinetic properties of a voltage-dependent junctional conductance. J Gen Physiol. 1981;77:95–117. doi: 10.1085/jgp.77.1.95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Bukauskas FF, Angele AB, Verselis VK, Bennett MVL. Coupling asymmetry of heterotypic connexin 45/ connexin 43-EGFP gap junctions: Properties of fast and slow gating mechanisms. Proc Natl Acad Sci USA. 2002;99:7113–7118. doi: 10.1073/pnas.032062099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Morley GE, Taffet SM, Delmar M. Intramolecular interactions mediate pH regulation of connexin43 channels. Biophys J. 1996;70:1294–1302. doi: 10.1016/S0006-3495(96)79686-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Moreno AP, Fishman GI, Beyer EC, Spray DC. Voltage dependent gating and single channel analysis of heterotypic gap junction channels formed of Cx45 and Cx43. In: Kanno Y, et al., editors. Intercellular Communication Through Gap Junctions, Progress in Cell Research. Vol. 4. Amsterdam: Elsevier Science B.V.; 1995. pp. 405–408. [Google Scholar]
  • 27.Bukauskas FF, Bukauskiene A, Bennett MVL, Verselis VK. Gating properties of gap junction channels assembled from connexin43 and connexin43 fused with green fluorescent protein. Biophys J. 2001;81:137–152. doi: 10.1016/S0006-3495(01)75687-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Bukauskas FF, et al. Clustering of connexin 43-enhanced green fluorescent protein gap junction channels and functional coupling in living cells. Proc Natl Acad Sci USA. 2000;97:2556–2561. doi: 10.1073/pnas.050588497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Müller DJ, Hand GM, Engel A, Sosinsky GE. Conformational changes in surface structures of isolated connexin 26 gap junctions. EMBO J. 2002;21:3598–3607. doi: 10.1093/emboj/cdf365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Lal R, John SA, Laird DW, Arnsdorf MF. Heart gap junction preparations reveal hemiplaques by atomic force microscopy. Am J Physiol. 1995;268:C968–C977. doi: 10.1152/ajpcell.1995.268.4.C968. [DOI] [PubMed] [Google Scholar]
  • 31.González D, Gómez-Hernández JM, Barrio LC. Molecular basis of voltage dependence of connexin channels: An integrative appraisal. Prog Biophys Mol Biol. 2007;94:66–106. doi: 10.1016/j.pbiomolbio.2007.03.007. [DOI] [PubMed] [Google Scholar]
  • 32.Bennett MVL, Verselis V, White RL, Spray DC. Gap junctional conductance: Gating. In: Hertzberg EL, Johnson RG, editors. Gap Junctions. New York: Alan R. Liss,Inc.; 1988. pp. 287–304. [Google Scholar]
  • 33.Komukai K, Brette F, Orchard CH. Electrophysiological response of rat atrial myocytes to acidosis. Am J Physiol Heart Circ Physiol. 2002;283:H715–H724. doi: 10.1152/ajpheart.01000.2001. [DOI] [PubMed] [Google Scholar]
  • 34.Délèze J. The recovery of resting potential and input resistance in sheep heart injured by knife or laser. J Physiol. 1970;208:547–562. doi: 10.1113/jphysiol.1970.sp009136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bukauskas F. Electrophysiology of the normal-to-hypoxic transition zone. Circ Res. 1982;51:321–329. doi: 10.1161/01.res.51.3.321. [DOI] [PubMed] [Google Scholar]
  • 36.Elliott AC, Smith GL, Eisner DA, Allen DG. Metabolic changes during ischaemia and their role in contractile failure in isolated ferret hearts. J Physiol. 1992;454:467–490. doi: 10.1113/jphysiol.1992.sp019274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Bountra C, Kaila K, Vaughan-Jones RD. Mechanism of rate-dependent pH changes in the sheep cardiac Purkinje fibre. J Physiol. 1988;406:483–501. doi: 10.1113/jphysiol.1988.sp017392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Garlick PB, Radda GK, Seeley PJ. Studies of acidosis in the ischaemic heart by phosphorus nuclear magnetic resonance. Biochem J. 1979;184:547–554. doi: 10.1042/bj1840547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Kreuzberg M, et al. Connexin 30.2 containing gap junction channels decelerate impulse propagation through the atrioventricular node. Proc Natl Acad Sci USA. 2006;108:5959–5964. doi: 10.1073/pnas.0508512103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Li X, Simard JM. Connexin45 gap junction channels in rat cerebral vascular smooth muscle cells. Am J Physiol Heart Circ Physiol. 2001;281:H1890–H1898. doi: 10.1152/ajpheart.2001.281.5.H1890. [DOI] [PubMed] [Google Scholar]
  • 41.Veruki ML, Hartveit E. Meclofenamic acid blocks electrical synapses of retinal AII amacrine and on-cone bipolar cells. J Neurophysiol. 2009;101:2339–2347. doi: 10.1152/jn.00112.2009. [DOI] [PubMed] [Google Scholar]
  • 42.Krüger O, et al. Defective vascular development in connexin 45-deficient mice. Development. 2000;127:4179–4193. doi: 10.1242/dev.127.19.4179. [DOI] [PubMed] [Google Scholar]
  • 43.Kumai M, et al. Loss of connexin45 causes a cushion defect in early cardiogenesis. Development. 2000;127:3501–3512. doi: 10.1242/dev.127.16.3501. [DOI] [PubMed] [Google Scholar]
  • 44.Allen DG, Xiao XH. Role of the cardiac Na+/H+ exchanger during ischemia and reperfusion. Cardiovasc Res. 2003;57:934–941. doi: 10.1016/s0008-6363(02)00836-2. [DOI] [PubMed] [Google Scholar]
  • 45.Bukauskas FF, Gutman AM, Kisunas KJ, Veteikis RP. Electrical cell coupling in rabbit sino atrial node and atrium. Experimental and theoretical evaluation. In: Bouman LN, Jongsma HJ, editors. Cardiac Rate and Rhythm. Physiological, Morphological and Developmental Aspects. The Hague: Martinus Nijhoff Publishers; 1982. pp. 195–216. [Google Scholar]
  • 46.Wojtczak J. Contractures and increase in internal longitudinal resistance of cow ventricular muscle induced by hypoxia. Circ Res. 1979;44:88–95. doi: 10.1161/01.res.44.1.88. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES