Abstract
The large-conductance Ca2+-activated K+ (BK) channels play an important role in the regulation of cellular excitability in response to changes in intracellular metabolic state and Ca2+ homeostasis. In vascular smooth muscle, BK channels are key determinants of vasoreactivity and vital-organ perfusion. Vascular BK channel functions are impaired in diabetes mellitus, but the mechanisms underlying such changes have not been examined in detail. We examined and compared the activities and kinetics of BK channels in coronary arterial smooth muscle cells from Lean control and Zucker Diabetic Fatty (ZDF) rats, using single-channel recording techniques. We found that BK channels in ZDF rats have impaired Ca2+ sensitivity, including an increased free Ca2+ concentration at half-maximal effect on channel activation, a reduced steepness of Ca2+ dose-dependent curve, altered Ca2+-dependent gating properties with decreased maximal open probability, and a shortened mean open-time and prolonged mean closed-time durations. In addition, the BK channel β-subunit-mediated activation by dehydrosoyasaponin-1 (DHS-1) was lost in cells from ZDF rats. Immunoblotting analysis confirmed a 2.1-fold decrease in BK channel β1-subunit expression in ZDF rats, compared with that of Lean rats. These abnormalities in BK channel gating lead to an increase in the energy barrier for channel activation, and may contribute to the development of vascular dysfunction and complications in type 2 diabetes mellitus.
INTRODUCTION
The large-conductance Ca2+-activated K+ (BK) channels are allosterically regulated by intracellular free Ca2+ and membrane potentials (1,2). The large conductance and high density in vascular smooth muscles make these channels a key determinant in the regulation of vascular tone (3). Functional vascular BK channels are composed of pore-forming α-subunits (encoded by the Slo gene) and accessory β1-subunits in a 4:4 stoichiometry (4,5). The α-subunit has seven transmembrane domains (S0–S6), including the highly conserved K+ channel selectivity filter between S5–S6, and a voltage sensor at S4. The C-terminus has four hydrophobic segments (S7–S10) that contain a number of important regulatory sites, including two regulators of conductance for potassium (RCK1 and RCK2) (6,7) and two high-affinity Ca2+-sensing regions, with Ca2+ concentration at half-maximal effect (EC50) in the 10−6 M range (8,9). One is the important high-affinity Ca2+ sensor, Ca2+ bowl (892-DQDDDDDPD-900) in the RCK2 domain; the other (D362/D367) is located in the RCK1 domain (10). A recent study proposed the presence of intersubunit interfaces that connect RCK1 and RCK2 between adjacent hSlo subunits through hydrophobic interactions, and four RCK1/RCK2 dimers may form an octameric ring in a homotetrameric hSlo channel (11). There is a third Ca2+ sensor in the RCK/Rossman folding domain (E374 and E399) that binds Mg2+ and Ca2+ at low affinity with EC50 in the 10−3 M range (12,13). The extracellular N-terminus of the α-subunit is thought to be important for functional coupling with the β-subunit (14). Recent reports also suggested that intracellular regions of the α-subunit, including S1, S2, and S3, are required for functional and physical interaction between the BK channel α-subunits and β-subunits (15,16). Coexpression of the β1-subunit significantly enhances the sensitivity of BK channel α-subunit to voltage-dependent and Ca2+-dependent activation, and modulates channel activation, channel deactivation (17–22), and single-channel kinetics (23). The physiological importance of the β1-subunit is demonstrated by studies with β1-gene knockout mice, which have increased vascular tone and hypertension with uncoupling of Ca2+ sparks to BK channels in vascular smooth muscle cells (24,25). Despite the important physiological functions rendered by the β-subunit, the role it plays in disease states such as diabetes mellitus has not been studied in detail.
Mortality associated with cardiovascular complications in diabetes mellitus has increased by 75% in the past decade (26), contributing to more than 200,000 deaths annually in the United States (http://www.diabetes.org). The cause of vascular dysfunction in diabetes is multifactorial and includes impaired endothelial function, reduced bioavailability of endothelium-derived relaxation factors (such as nitric oxide, prostaglandin I2, and endothelium-derived hyperpolarizing factors), and enhanced sensitivity to endothelium-derived constricting factors (27) and angiotensin II (28). In vitro vasoreactivity studies showed that diabetic vascular dysfunction is associated with impaired K+ channel activities in smooth muscle cells, including those of the voltage-gated K+ (Kv) channels (29), ATP-sensitive K+ (KATP) channels (30–32), small-conductance Ca2+-activated K+ (SK) channels (33), and BK channels (33–37). Coronary artery smooth muscle cells cultured with high glucose showed suppressed Kv channel activation by enhanced production of hydrogen peroxide (H2O2) and peroxynitrite, which directly inhibit channel activity (38) and impair cAMP-dependent signaling regulation (29,39). The mechanisms that underlie BK channel dysfunction in diabetic vessels are dependent on the animal model, vessel bed, and stage of diabetes. In high-fructose diet-induced insulin-resistant rats, BK current density in mesenteric artery smooth muscle cells was reduced, but the channel Ca2+sensitivity and voltage sensitivity were unchanged (37). In streptozotocin (STZ)-induced type I diabetic rats, the down-regulation of BK channel activity and decreased sensitivity to voltage-dependent activation were associated with impaired β1-subunit function (35). During the early stages of diabetes in Zucker Diabetic Fatty (ZDF) rats, a genetically determined type 2 diabetic animal model, reduced BK channel participation in arachidonic acid-mediated coronary vasodilatation was mostly attributable to impaired biosynthesis of prostaglandin I2 (36). In ZDF rats with more advanced diabetes, the BK channel response to Ca2+ and NS1619 (a BK channel α-subunit activator) activation was reduced, suggesting that the function of the α-subunit was impaired (33). Tang et al. found that the cysteine residue at 911 (C911) of hSlo, which is close to the Ca2+ bowl, is the major functional target of redox modulation by H2O2 (40). We also reported that high glucose-induced accumulation of reactive oxygen species suppressed the current density, and slowed the activation and deactivation of hSlo expressed in HEK293 cells (41). However, most of these previous studies focused on physiological regulation and on the role of BK channels in diabetic pathophysiology. The changes in biophysical properties of BK channels, such as channel kinetics and gating behaviors in diabetic vascular smooth muscle cells, have not been characterized in detail. In this study, we hypothesized that the intrinsic gating properties of vascular BK channels were altered in diabetes. Using single-channel recording techniques, we compared BK channel activities and channel protein expression in vascular smooth muscle cells from ZDF rats and age-matched Lean control rats. Our results showed profound abnormalities in the regulation of BK channels of ZDF rats. In particular, Ca2+-dependent gating behaviors are substantially altered with the loss of β1-subunit-mediated channel activation, and such changes can be explained by a significant reduction of BK channel β1-subunit expression in ZDF rats. Finally, we employed a simple model to describe and better understand the differences in BK channel kinetics between Lean and ZDF rats. Our results may provide important insights into the fundamental mechanisms that underlie the development of vascular dysfunction and complications in diabetes.
MATERIALS AND METHODS
Animals
Male ZDF rats (Gmi-fa/fa) and age-matched Lean control rats (Gmi-fa/+ or +/+) were obtained from Charles River Laboratories (Wilmington, MA) at 6–8 weeks of age. All rats were housed in the Animal Care Facility of Mayo Clinic, and received a Purina 3008 modified mouse/rat diet, according to a protocol approved by the Animal Care and Use Committee of the Mayo Foundation. Blood was drawn from the tail vein, and blood glucose was monitored using an Accu-Chek glucose meter (Roche Diagnostics Co., Indianapolis, IN). Rats with blood glucose levels consistently higher than 300 mg/dL were considered diabetic. The ZDF rats 4–6 months after the development of hyperglycemia and age-matched Lean control rats were sacrificed for the experiments.
Isolation of coronary arterial smooth muscle cells
Coronary artery smooth muscle cells were dissociated enzymatically, as previously reported (42). Briefly, rat hearts were rapidly excised and placed in cold (4°C) physiological saline solution that contained 145.0 mM NaCl, 4.0 mM KCl, 0.05 mM CaCl2, 1.0 mM MgCl2, 10.0 mM HEPES, and 10.0 mM glucose (pH 7.2). The septal, right, and left anterior descending coronary arteries were carefully dissected free of surrounding myocardium and connective tissue, and placed in 1 mL physiological saline solution containing bovine serum albumin (0.1%, w/v) for a 10-min incubation at 37°C in a shaking water bath. The vessels were treated with 1.75 mg papain and 1.25 mg dithiothreitol in 1.0 mL saline solution for 10 min, and further digested with 1.25 mg collagenase and 1.25 mg trypsin inhibitor in 1.0 mL saline solution at 37°C for 10 min. The vessels were then washed three times with 1.0-mL aliquots of saline solution, and gently triturated with a fire-polished glass pipette until cells were completely dissociated.
Inside-out single-channel recordings
Single BK currents from coronary artery smooth muscle cells were recorded in the inside-out configuration, using an Axopatch 200B integrating amplifier and pCLAMP 8.2 software (Axon Instruments, Foster City, CA). The output signals were filtered with an eight-pole Bessel filter (902 LPF, Frequency Devices, Inc., Haverhill, MA) at 5 kHz, and digitized at 20 kHz. Patch pipettes had a typical tip resistance of 5–10 MΩ when filled with the pipette solution, which contained 140.0 mM KCl, 1.0 mM CaCl2, 1.0 mM MgCl2, 10.0 mM HEPES, and 1.0 mM EGTA (pH 7.4 with KOH). The bath solution contained 140.0 mM KCl, 1.0 mM MgCl2, 1.0 mM EGTA, and 10.0 mM HEPES (pH 7.35 with KOH). Various amounts of Ca2+ were added in the bath solution to obtain the desired free Ca2+ concentrations (from 10−9 to 10−4 M), calculated using Chelator software, as previously described (43). The number of channels in the excised patch was determined at +60 mV in the presence of 100 μM free Ca2+, a condition that was shown to activate BK channel openings maximally in rat coronary artery smooth muscle cells (43). The BK channels were identified by their unitary conductance, voltage-sensitivity, and Ca2+-sensitivity. Open probability (pO) in patches with multiple channels was determined using Fetchan in pClamp 8.2, based on the equation:
![]() |
where T is the duration of recording, tj is the time spent with j = 1,2,3, … n channel openings, and N is the maximal number of channel openings observed when pO was high.
Because FETCHAN, the single-channel analysis program in pCLAMP 8.2, only allows a maximum presence of five channels for analysis, excised patches containing five or fewer channels were used for all pO measurements.
Single-channel kinetic analysis
Excised patches containing only a single channel were used for kinetics analysis. The digital current signals were filtered at a bandwidth of 2.0 kHz with a digital Gaussian filter, and events were detected by the half-amplitude threshold criterion, using TAC software (Bruxton, Inc., Seattle, WA) with a dead time of 90 μs. The effects of the Gaussian filter on channel dwell-time durations at 50% threshold were corrected by cubic spline interpolation of the digital signal, as previously described (44). The events of subconductance opening were excluded for kinetics analysis. Dwell-time histograms were fitted with the sums of exponential probability density functions, using the TACFit program. The number of exponential components was determined by the maximal likelihood ratio test, and additional exponential components were included only when the p > 0.95. In the TACFit program, burst resolution was given by the burst opening that is defined by intraburst gaps shorter than the critical closed time (tcrit), calculated as 1.0 ms in our recording system, using the following equation (44):
![]() |
where τf and αf represent the fast time constant and its relative weight, respectively, and τs and αs represent the slow time constant and its relative weight, respectively.
Single-channel kinetics analysis was performed on results obtained over a range of intracellular Ca2+ concentrations, using MIL software (QuB Suite, State University of New York at Buffalo, Buffalo, NY) that employs logarithm likelihood ratio tests, corrects the missed events, and gives error estimates of the fitted parameters (45,46).
The Ca2+-dependent channel activation was fitted using the Hill equation (43):
![]() |
where pO,max is the maximal channel open probability, [Ca2+] represents the intracellular free Ca2+ concentration, EC50 is the concentration at half-maximal effect, and nH is the Hill coefficient.
The pO − voltage (V) relationships were characterized by the Boltzmann equation (43):
![]() |
where V1/2 is the voltage at which pO is half of pO,max, V is the membrane potential, z is the number of equivalent charge movements, e is the elementary charge, F is the Faraday constant, R is the universal gas constant, and T is the absolute temperature.
The change in free energy of Ca2+-binding contributions to channel activation was calculated using the equation ΔΔCa = −Δ(zeV1/2), based on the pO − V curve shift in response to changes of the free Ca2+ concentrations (13). Curve fittings were performed using Igor 6.01 software (WaveMetrics, Inc., Lake Oswego, OR).
Western-blot analysis
Western blots were performed as previously described (36). Isolated aortas from three pairs of Lean and ZDF rats were homogenized, electrophoresed, transferred to nitrocellulose membrane, and immunoblotted against rabbit anti-human BK channel antibodies (1:200, custom-made) and rat anti-sloβ1 antibodies (1:200, Alomona Labs, Jerusalem, Israel). Blots were also probed with anti-β-actin antibodies (1:500, Santa Cruz, CA) as loading controls. After extensive washing, horseradish peroxidase-conjugated secondary antibodies were added. Signals were developed using an Immun-Star HRP Chemiluminescent Kit (Bio-Rad, Hercules, CA). Optical density of the bands was analyzed using Scion Image software (Scion, Frederick, MD). Protein expression was expressed as relative abundance normalized to β-actin.
Chemicals
Dehydrosoyasaponin-1 (DHS-1) was a gift from Merck Research Laboratories (Merck & Co., Inc., Rahway, NJ). All other chemicals were purchased from Sigma-Aldrich Co. (St. Louis, MO).
Statistical analysis
Data are presented as mean ± SE. Student's t-test was used to compare data between two groups. A paired t-test was used to compare data before and after treatment. A statistically significant difference was defined as p < 0.05.
RESULTS
Development of hyperglycemia in ZDF rats
Blood glucose in ZDF rats started to increase after 8 weeks of age. At the time of the experiment, the Lean and ZDF rats were 27.3 ± 5.7 weeks of age (n = 14) and 29.8 ± 1.8 weeks of age (n = 14, vs. Lean rats, p = NS), respectively. The ZDF rats had markedly elevated blood glucose (544.8 ± 19.1 mg/dL vs. 179.2 ± 10.1 mg/dL in Lean rats, n = 14, p < 0.05), but body weights were not significantly different from Lean rats (429.6 ± 5.7 g in Lean rats vs. 395.8 ± 21.2 g in ZDF rats, n = 14 for both, p = NS), similar to previous reports (33).
Impaired Ca2+-dependent BK channel activation in coronary artery smooth muscle cells in ZDF rats
We examined the Ca2+-sensitivity of BK channels in Lean and ZDF rats. Fig. 1 A shows Ca2+-dependent BK channel activities elicited at +60 mV from coronary artery smooth muscle cells of Lean and ZDF rats in the presence of various free Ca2+ concentrations. There was no spontaneous channel opening when the membrane was excised into a bath solution that contained 0.01 μM free Ca2+. The BK channel activities in both Lean and ZDF rats became more robust with an increase of free Ca2+ in a dose-dependent manner. The BK channel pO between Lean and ZDF rats became significantly different at Ca2+ concentrations higher than 0.5 μM. In Lean rats, pO was 0.54 ± 0.08 at 1 μM (n = 9), 0.76 ± 0.07 at 10 μM (n = 9), and 0.78 ± 0.05 at 100 μM Ca2+ (n = 9), and pO was significantly reduced in ZDF rats to 0.27 ± 0.07 at 1 μM (n = 11, p < 0.05 vs. Lean), 0.58 ± 0.06 at 10 μM (n = 11, p < 0.05 vs. Lean), and 0.63 ± 0.04 at 100 μM Ca2+ (n = 11, p < 0.05 vs. Lean). The Ca2+ dose-dependent curve was fitted with an EC50 of 1.03 ± 0.16 μM in Lean control rats (n = 9) and 2.44 ± 0.48 μM in ZDF rats (n = 11 vs. Lean, p < 0.05). Hence, the BK channel pO − Ca2+ relationship shifted rightward and downward in ZDF rats (Fig. 1 B). Most importantly, nH was reduced from 4.1 in Lean rats to 1.1 in ZDF rats, indicating that cooperativity of Ca2+-dependent channel activation was absent in ZDF rats.
FIGURE 1.
BK channels of coronary artery smooth muscle cells in ZDF rats have reduced Ca2+ sensitivity. (A) Representative inside-out single BK channel currents were recorded at +60 mV from Lean rats (left column) and ZDF rats (right column) in the presence of various free Ca2+ concentrations (0.01 μM to 100 μM). BK currents were activated by an increase in free Ca2+ concentration in both Lean and ZDF rats. (B) pO − [Ca2+] relationships in Lean (n = 9) and ZDF (n = 11) rats were fitted using the Hill equation. There was a significant reduction in pO in ZDF rats at Ca2+ concentrations ≥1 μM. Data are presented as mean ± SE. *p < 0.05 vs. Lean. Here and in subsequent figures, C and O represent closed and open states of the BK channel, respectively.
Decreased free energy of Ca2+-dependent channel activation in coronary smooth muscle cells in ZDF rats
Fig. 2 A shows representative BK channel tracings of Lean and ZDF rats recorded at different membrane potentials in the presence of 1 μM free Ca2+. The BK channel activity increased with membrane potential depolarization in Lean and ZDF rats, both of which had the same unitary current amplitudes at the same given voltage. No spontaneous BK channel opening was discernible at −60 mV in either Lean or ZDF rats. Occasional channel openings began to appear at −20 mV in Lean rats and at +20 mV in ZDF rats (Fig. 2 A, left column), suggesting a rightward shift in the threshold of channel voltage-dependent activation in ZDF rats. Indeed, at all tested voltages above 0 mV, channel openings were more robust in Lean rats, and there was a significant reduction in channel pO in ZDF rats over the full range of voltages, from +20 mV (0.06 ± 0.03 in ZDF rats, n = 13, vs. 0.31 ± 0.09 in Lean rats, n = 15, p < 0.05) to +120 mV (0.59 ± 0.06 in ZDF rats, n = 15, vs. 0.71 ± 0.06 in Lean rats, n = 13, p < 0.05). Fig. 2 B shows the BK channel pO − V relationships of Lean and ZDF rats in the absence of free Ca2+, and there was a reduced maximal pO in ZDF rats, measured at voltages higher than 150 mV (0.48 ± 0.06 in Lean rats vs. 0.20 ± 0.07 in ZDF rats, p < 0.05, n = 5 for both). There was no significant difference between Lean and ZDF rats in the voltage at half-maximal activation (V1/2) (144.9 ± 8.8 mV in Lean rats, n = 5, and 132.6 ±10.8 mV in ZDF rats, n = 5, p = NS) or in the equivalent charge movement z (0.8 e in Lean rats and 0.7 e in ZDF rats). In the presence of 1 μM Ca2+, V1/2 was 41.3 ± 4.5 mV (n = 13) in Lean rats, and V1/2 shifted to the right in ZDF rats, at 89.6 ± 6.5 mV (n = 15, p < 0.05 vs. Lean rats), whereas z was unchanged, at 1.0 e in Lean rats and 1.1 e in ZDF rats (Fig. 2 C). Thus, the change in the intrinsic free energy of Ca2+-binding that contributes to channel activation can be estimated, based on the shift of pO − V relationships between Lean and ZDF rats: ΔΔCa= −Δ(zeV1/2) (13). The ΔΔCa was −41.4 ± 6.7 (n = 5) kJ/mol in Lean rats, but was reduced by 62.3% to −25.8 ± 8.7 (n = 5, p < 0.05) in ZDF rats, when free Ca2+ was increased from 0 μM to 1 μM, suggesting that Ca2+-dependent activation of BK channels is less favorable in ZDF rats compared with Lean controls.
FIGURE 2.
BK channels in ZDF rats have reduced response to voltage activation. (A) Representative tracings of single BK channel activities recorded in presence of 1 μM free Ca2+ at various voltages (from −60 mV to +100 mV). Channel activity was enhanced by membrane depolarization, with an activation threshold of −20 mV in Lean rats (left column) and +20 mV in ZDF rats (right column). (B) The pO − V relationships in Lean (n = 15) and ZDF (n = 13) rats in the absence of Ca2+. The pO was significantly lower in ZDF rats at voltages higher than +150 mV. (C) The pO − V relationships in Lean (n = 15) and ZDF (n = 13) rats in the presence of 1 μM Ca2+. The pO was significantly lower in ZDF rats at voltages between +20 and +120 mV, with a rightward and downward shift of the pO − V curve. Data are presented as mean ± SE. *p < 0.05 vs. Lean.
Altered BK channel kinetics in ZDF rats
For a better understanding of altered Ca2+-dependent BK channel activation in ZDF rats, we further examined the Ca2+-dependent gating properties in both Lean and ZDF rats. Because BK channels of ZDF rats were almost inactive at Ca2+ concentrations <0.1 μM, we compared single-channel gating between Lean and ZDF rats at Ca2+ concentrations from 1 μM to 100 μM, with a testing potential of +60 mV. Fig. 3 A illustrates typical tracings of inside-out single-channel BK currents in Lean and ZDF rats with expanded details. In the presence of 1 μM Ca2+, BK channel pO was much higher in Lean rats than in ZDF rats. Few very brief subconductance openings were evident in both Lean and ZDF rats. An increase of Ca2+ to 100 μM markedly increased the channel pO (>0.8) in both Lean and ZDF rats, but more frequent and prolonged subconductance openings were observed in ZDF rats (Fig. 3 B). The effects of Ca2+ on mean open times and on mean closed durations of BK channels in Lean and ZDF rats are illustrated in Fig. 3, C and D, respectively. Compared with Lean controls, ZDF rats had mean open times that were markedly shortened at Ca2+ concentrations ≥1 μM, and mean closed durations that were significantly prolonged at Ca2+ concentrations ≥1 μM. Because normal intracellular Ca2+ concentration can reach >10 μM concentrations, especially in microdomains where calcium sparks are elicited (47,48), these fundamental changes in BK channel properties are physiologically relevant. These results indicate that there is a reduction in BK channel sensitivity to activation by Ca2+ in ZDF rats.
FIGURE 3.
BK channels in ZDF rats have shortened mean burst durations and prolonged mean closed durations. Representative tracings of single BK currents were elicited at +60 mV in the presence of 1 μM free Ca2+ (A) and 100 μM free Ca2+ (B), in Lean (upper tracings) and ZDF (lower tracings) rats. Selected tracing segments with expanded details show longer channel-opening durations and higher pO in Lean rats, and more frequent and stabilized subconductance levels in ZDF rats. Note that four levels of subconductance were discerned, represented by Sub1, Sub2, Sub3, and Sub4 (fully open). Plots of relationships between Ca2+ concentrations and mean burst durations (C) and mean closed times (D) of BK channels in Lean and ZDF rats are shown. Compared with Lean rats, ZDF rats had shorter mean burst durations and longer mean closed durations. Data are presented as mean ± SE. *p < 0.05 vs. Lean (n = 6).
Representative histograms of the BK channel probability density function in the presence of 1 μM, 10 μM, and 100 μM Ca2+ in Lean and ZDF rats are given in Fig. 4, A and B, respectively. For the best fit, the open dwell-time distribution histograms required at least three components: fast (τO1), intermediate (τO2), and slow (τO3) open-time constants. The closed dwell-time histograms had four components: fast (τC1), intermediate (τC2), slow (τC3), and very slow (τC4) closed-time constants. In Lean rats, an increase in Ca2+ concentration prolonged the open dwell times and abbreviated the closed dwell times. However, such Ca2+-dependent relationships were not closely followed in ZDF rats (Table 1). For example, an increase in Ca2+ markedly prolonged all three open-time constants, τO1, τO2, and τO3, in Lean rats. In ZDF rats, an increase in Ca2+ prolonged τO1 and τO2, but τO3 was shortened. An increase in Ca2+ shortened the BK-channel closed-time constants in Lean and ZDF rats, but at any given Ca2+ concentration, τC1, τC2, τC3, and τC4 were significantly lengthened in ZDF rats compared with Lean rats. A comparison of BK channel open-time and closed-time constants between Lean and ZDF rats at Ca2+ concentrations of 1 μM, 10 μM, and 100 μM is summarized in Table 1.
FIGURE 4.
Single BK channel kinetics in Lean and ZDF rats. Representative histograms of single BK channel open and closed dwell-time durations (from one patch) in the presence of 1 μM, 10 μM, and 100 μM free Ca2+ in Lean rats (A) and ZDF rats (B). Data were obtained from an excised patch in inside-out configuration. Dwell-time distributions were best described by three open dwell-time components and four closed dwell-time components. An increase in cytoplasmic free Ca2+ from 1 μM to 100 μM prolonged all open dwell-time constants (τO), and shortened all closed dwell time constants (τC). Dashed lines represent distribution of exponential components, determined by the logarithm likelihood ratio test. The values of each time constant component and its relative weight (in parentheses) are given above each histogram.
TABLE 1.
BK channel open and closed dwell-time constants and their relative weights in Lean and ZDF rats
| Time constant (weight) | 1 μM, Lean | 1 μM, ZDF | 10 μM, Lean | 10 μM, ZDF | 100 μM, Lean | 100 μM, ZDF |
|---|---|---|---|---|---|---|
| τO1 (ms) | 0.55 ± 0.06 | 0.48 ± 0.06 | 0.85 ± 0.25 | 0.63 ± 0.03 | 0.74 ± 0.07 | 0.31 ± 0.19 |
| (AO1) | (0.40 ± 0.07) | (0.26 ± 0.03) | (0.35 ± 0.08) | (0.47 ± 0.08) | (0.54 ± 0.05) | (0.36 ± 0.06)* |
| τO2 (ms) | 9.63 ± 1.95 | 4.77 ± 1.06* | 10.81 ± 3.55 | 6.08 ± 1.28 | 13.92 ± 4.04 | 10.74 ± 4.15 |
| (AO2) | (0.23 ± 0.04) | (0.41 ± 0.06)* | (0.21 ± 0.04) | (0.39 ± 0.07)* | (0.26 ± 0.06) | (0.34 ± 0.01)* |
| τO3 (ms) | 43.23 ± 9.33 | 39.73 ± 8.40 | 61.35 ± 10.70 | 22.47 ± 4.05* | 61.97 ± 8.60 | 20.07 ± 3.30* |
| (AO3) | (0.38 ± 0.09) | (0.33 ± 0.07) | (0.332 ± 0.11) | (0.26 ± 0.07) | (0.22 ± 0.08) | (0.30 ± 0.09) |
| τC1 (ms) | 0.39 ± 0.06 | 0.61 ± 0.10 | 0.49 ± 0.06 | 0.58 ± 0.13 | 0.46 ± 0.06 | 0.54 ± 0.13 |
| (AC1) | (0.61 ± 0.03) | (0.58 ± 0.09) | (0.62 ± 0.05) | (0.61 ± 0.09) | (0.61 ± 0.03) | (0.69 ± 0.09) |
| τC2 (ms) | 2.51 ± 0.472 | 4.26 ± 0.75 | 2.26 ± 0.42 | 4.48 ± 1.40 | 2.33 ± 0.61 | 4.74 ± 0.51* |
| (AC2) | (0.24 ± 0.03) | (0.20 ± 0.02) | (0.26 ± 0.02) | (0.27 ± 0.05) | (0.29 ± 0.01) | (0.19 ± 0.03)* |
| τC3 (ms) | 29.27 ± 10.41 | 198.59 ± 48.26* | 29.62 ± 10.46 | 92.43 ± 26.17* | 20.73 ± 9.78 | 74.02 ± 22.01 |
| (AC3) | (0.03 ± 0.02) | (0.19 ± 0.07) | (0.07 ± 0.02) | (0.14 ± 0.07) | (0.11 ± 0.03) | (0.12 ± 0.08) |
| τC4 (ms) | 165.2 ± 37.74 | 3447.2 ± 910.9* | 93.67 ± 17.38 | 1706.2 ± 701.6* | 75.84 ± 11.88 | 852.18 ± 59.29* |
| (AC4) | (0.03 ± 0.02) | (0.04 ± 0.01) | (0.06 ± 0.02) | 0.05 ± 0.04 | (0.05 ± 0.02) | (0.09 ± 0.05) |
τO1, τO2, and τO3 represent fast, intermediate, and slow components of channel open dwell-time durations. τc1, τc2, τc3, and τc4 represent fast, intermediate, slow, and very slow components of channel closed dwell-time durations, respectively. AO1–3 and AC1–4 represent the relative weight of each open and closed dwell-time constant, where AO1 + AO2 + AO3 = 1, and AC1 + AC2 + AC3 + AC4 = 1. Data are represented as mean ± SE, n = 6.
p < 0.05 vs. Lean.
It was shown that each BK channel α-subunit has three Ca2+ binding sites. We demonstrate that there is a reduced nH for Ca2+-dependent BK channel activation in ZDF rats, suggesting either a reduction in the number of Ca2+ binding sites or in the cooperativity of Ca2+ binding (Fig. 1 B). To describe completely all the features of BK channel Ca2+-dependent and voltage-dependent activation requires a 50-state, two-tiered model (2,49). Here, we tried to use the simplest model that could describe our experimental results. Because BK channel kinetics analysis suggests the presence of at least three open states and four closed states in the range of Ca2+ concentrations tested, a minimal kinetic scheme could be represented by Scheme I in Fig. 5 A (50), where the transitions between states C0↔C1, C1↔C2, and C2↔C3 are Ca2+-dependent. Using single-channel kinetic-analysis software, we estimated the rate constants between each state transition of the BK channel in Lean and ZDF rats, and the group results from three experiments are listed in Table 2. The energy barrier for channel activation could be calculated using the equation G = RT lnKeq (51), where G is the energy barrier, Keq represents the equilibrium constant of the transition state, R is the universal gas constant, and T is the absolute temperature. Hence, the total energy barriers (GT) can be determined by the sum of each energy barrier of the transition state. The relationships between GT and the logarithm of Ca2+ concentrations are plotted in Fig. 5 B (n = 3 for Lean and ZDF rats, respectively). The GT − log[Ca2+] relationships can be fitted by a linear equation with a slope of 16.6 for both Lean and ZDF rats. However, the GT for BK channel activation was increased by 8.6-fold at all given Ca2+ concentrations. The upward shift of the GT − log[Ca2+] relationship in ZDF rats indicates less thermodynamically favorable conditions of Ca2+-dependent BK channel activation in ZDF rats, and that at any given Ca2+ concentration, there is a greater energy barrier to overcome for BK channel activation. We did not include the transitions between O1↔O2 and O2↔O3 in the model, because the loop transition model would not change the values of the total energy barrier obtained from Scheme I (Fig. 5 A), and would not change our conclusions.
FIGURE 5.
Increased energy barriers of Ca2+-dependent BK channel activation in ZDF rats. (A) BK channel-gating can be described by a simple kinetics model, as shown in Scheme I. O and C represent the channel open states and closed states, where transitions between C0↔C1, C1↔C2, and C2↔C3 are Ca2+-dependent. (B) Total energy barriers (GT) for BK channel activation, calculated from Scheme I (Fig. 5 A) in Lean and ZDF rats, are plotted against the logarithm of Ca2+ concentrations. The GT − log[Ca2+] relationship is fitted with a linear equation, with GT = 2.5 + 16.6 log[Ca2+] in Lean rats, and GT = 21.4 + 16.6 log[Ca2+] in ZDF rats. Hence, an 8.6-fold increase in energy barrier for BK-channel activation is present in ZDF rats at all given Ca2+ concentrations.
TABLE 2.
State transition rate constants of BK channels in Lean and ZDF rats
| (S−1) | C0→C1* | C1→C0 | C1→C2* | C2→C1 | C2→C3* | C3→C2 | C1→O0 | O1→C1 | C2→O2 | O2→C2 | C3→O3 | O3→C3 |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Lean | 45.1 ± 23.5 | 5.8 ± 5.0 | 296 ± 239 | 1.0 ± 0.7 | 7.0 ± 6.6 | 218 ± 66 | 4.7 ± 2.9 | 167 ± 53 | 655 ± 170 | 202 ± 129 | 327 ± 138 | 202 ± 37.1 |
| ZDF | 0.9 ± 0.5 | 70.1 ± 69.7 | 23.2 ± 17.0 | 560 ± 210 | 0.9 ± 0.9 | 6.9 ± 6.8 | 2864 ± 526 | 128 ± 79 | 251 ± 21 | 2117 ± 613 | 17.8 ± 17.2 | 14.8 ± 9.4 |
Data are presented as mean ± SE (n = 3 for both groups).
Transition is Ca2+-dependent (in μM s−1).
Impaired BK channel activity with loss of β1-subunit-mediated channel activation and reduced β1-subunit expression in smooth muscle cells of ZDF rats
Because the BK channel β-subunit is known to regulate channel sensitivity to Ca2+, we examined the function of the β1-subunit using a specific BK channel β-subunit activator, DHS-1 (52), to determine whether β1-subunit function contributes to the mechanism underlying impaired BK channel function in ZDF rats. Fig. 6 A shows representative tracings of inside-out BK currents elicited at +60 mV in the presence of 0.5 μM free Ca2+, before and after exposure to 0.1 μM DHS-1, followed by a washing out of DHS-1. In Lean rats, DHS-1 significantly enhanced channel pO from 0.21 ± 0.09 at baseline to 0.54 ± 0.08 with DHS-1 (n = 7, p < 0.05 vs. baseline), and this effect was reversible upon washing out. In contrast, DHS-1 had no significant effects on BK channel activities in ZDF rats (0.17 ± 0.06 at baseline and 0.28 ± 0.08 with DHS-1, n = 7, p = NS), suggesting that β1-subunit-mediated BK channel activation was significantly reduced in ZDF rats. A comparison of the DHS-1 effects between Lean and ZDF rats is summarized in the bar graphs of Fig. 6 A. We then determined the protein expression of BK channel α-subunits and β1-subunits in the aortas of Lean and ZDF rats. The immunoblots of BK channel α-subunits and β1-subunits from three pairs of Lean and ZDF rat aortas are illustrated in Fig. 6 B. The group results of densitometric analysis are summarized in the bar graphs. There was no significant difference in BK channel α-subunit expression between Lean and ZDF rats (0.39 ± 0.02 in Lean rats vs. 0.37 ± 0.01 in ZDF rats, n = 3, p = NS). These results are similar to those reported by other laboratories (34). However, BK channel β1-subunit expression was reduced by 2.1-fold in ZDF rats, from 0.132 ± 0.03 in Lean rats to 0.062 ± 0.01 in ZDF rats (n = 3, p < 0.05). These important results indicate that β1-subunit expression is reduced, and that β1-subunit-mediated regulation of BK channel function is impaired, in ZDF rats.
FIGURE 6.
Impaired β1-subunit-dependent activation and reduced β1-subunit expression in BK channels of ZDF rats. (A) BK currents in Lean (left) and ZDF (right) rats were recorded from inside-out excised patches at +60 mV in presence of 0.5 μM free Ca2+ at baseline, with application of 0.1 μM DHS-1, followed by washing out. Application of DHS-1 significantly increased BK channel pO in Lean rats, but not in ZDF rats. Comparison of DHS-1 effects on BK channel activity between Lean and ZDF rats is summarized in bar graphs. (B) Immunoblot analysis of BK channel α-subunit and β1-subunit expression in aortas from three pairs of Lean and ZDF rats. Protein expression was expressed as relative abundance, normalized to β-action. There was a 2.1-fold decrease in β1-subunit expression in aortas of ZDF rats, whereas α-subunit expression was unchanged compared with Lean rats.
Increased BK channel subconductance openings in ZDF rats
It is known that BK channels exhibit multiple levels of subconductance, in which some but not all four channel subunits are in the open state and BK channel subconductance activity is thought to be regulated by the β1-subunit (53–56). We found that BK channel subconductance openings appeared to be more frequent and stable in ZDF rats, especially in the presence of high free Ca2+ (≥10 μM). Fig. 7 A illustrates a typical tracing recorded at +60 mV, with expanded details showing multiple subconductance levels. The amplitude histograms were fitted with a Gaussian function, and showed four levels of conductance (three levels of subconductance), i.e., Sub1, Sub2, Sub3, and Sub4 (fully open), with unitary current amplitudes of 4 pA, 8 pA, 12 pA, and 16 pA, respectively (Fig. 7 B). The cooperative conformational changes of channel subunits can be estimated from the relationship between the number of subconductance states and their relative frequencies. Relative frequencies were plotted against each subconductance state in Lean and ZDF rats in Fig. 7 C. These relationships can be fitted by a single exponential function: y = ω · exp(ψ · x), where ω is the fitting constant, and ψ is the coefficient of subunit conformational change. The coefficient of BK channel subunit conformational change was estimated to be 3.3 in Lean rats, and 1.4 in ZDF rats. These findings match our results, i.e., that the nH of Ca2+-dependent channel activation was reduced from 4.1 in Lean rats to 1.1 in ZDF rats (Fig. 1 B). Hence, nH in the Ca2+-dependent channel activation curve is in agreement with the cooperativity of subunit conformational change in tetrameric BK channels. These observations suggest that Ca2+ cooperativity and subunit conformational changes may be coupled, but such coupling appeared to be impaired in the BK channels of ZDF rats.
FIGURE 7.
Increased BK channel subconductance openings in ZDF rats. (A) Representative single BK current recordings were obtained at +60 mV in the presence of 1 μM free Ca2+, with selected segments that showed expanded details, demonstrating presence of four sublevels of openings. (B) Amplitude histogram was fitted using a Gaussian function, and showed four peaks with unitary amplitudes of 4 pA (Sub1), 8 pA (Sub2), 12 pA (Sub3), and 16 pS (Sub4 or fully open). Relative frequencies of each subconductance state were calculated by the area under each component of the Gaussian function. (C) Relative frequencies were plotted against subconductance states, and relationships were fitted using a single exponential function. Coefficient of subconductance conformational changes was estimated to be 3.3 in Lean rats (n = 3), and 1.4 in ZDF rats (n = 3), for a reduction of 2.3-fold in the latter.
DISCUSSION
We compared the single-channel kinetics of vascular BK channels between genetically determined type II diabetic ZDF rats and aged-matched Lean control rats. We made several novel observations. First, BK channels in ZDF rats have reduced Ca2+ sensitivity. Second, BK channels in ZDF rats have altered Ca2+-dependent gating, showing shortened open dwell times, prolonged closed dwell times, and more frequent subconductance openings. Third, BK channels in ZDF rats have a significant impairment in β1-subunit-mediated activation of BK channels, associated with a substantial decrease in BK channel β1-subunit expression but not in α-subunit expression. These alterations in the biophysical and biochemical properties of BK channels result in an increase of energy barriers for channel activation, and may contribute to diabetic vascular dysfunction.
It is well-known that the BK channel β1-subunit is an important regulator of BK channel function. It modulates the Ca2+-dependent activation and deactivation gating of whole-cell Slo currents (17–21), reduces the steepness of the Ca2+ response curve (21), increases the maximal pO, prolongs the open dwell times, and abbreviates the closed dwell times in single-channel kinetics (23). However, all these experiments were performed on expressed Slo channels, and the physiological relevance of the β1-subunit modulation of native BK channels is not clear. In this study, we provide compelling evidence that BK channels in native vascular smooth muscle cells from type II diabetic rats show altered gating behaviors that can be attributed to the deficiency of β1-subunit functions. We found that the cooperativity of Ca2+-dependent BK channel activation was lost in diabetes, as indicated by a reduction in the coefficient of subunit conformation and in the Hill coefficient. Moreover, kinetic analysis showed that BK channels in ZDF rats had a prolongation of closed dwell-time durations and a shortening of open dwell-time durations, with an increase in energy barriers for Ca2+-dependent BK channel activation. However, we do not know whether reduced Ca2+ cooperativity occurs among Ca2+ sensors of the same channel α-subunit or among those of different α-subunits of the tetrameric channel. Qian et al. (57) reported that intrasubunit cooperation is more effective in mediating channel Ca2+ sensitivity. Impairment of intrasubunit cooperativity reduces not only the efficiency but also the efficacy of Ca2+-dependent activation of hSlo, whereas impairment of intersubunit cooperativity affects only the efficiency of Ca2+-dependent activation (57). Because both the efficiency and efficacy of Ca2+-dependent activation were reduced in ZDF rats, it is likely that impaired intrasubunit Ca2+ cooperative binding is the culprit. The ZDF rats showed a rightward and downward shift in the pO − V curve, suggesting impaired voltage-dependent BK channel activation. Because the free energy of Ca2+-binding for channel activation was reduced but not the equivalent charge movement z, the voltage-dependent changes in ZDF rats are likely secondary to Ca2+-dependent changes in channel-gating properties.
A key finding in this study is that the function of BK channel β1-subunits is impaired in ZDF rats. Recently, in STZ-induced type I diabetic animals, BK channel β1-subunit mRNA transcripts were found to be reduced by 60% (35). However, using immunohistochemistry techniques, Burnham et al. reported that the expression of BK channel α-subunits and β1-subunits in the mesenteric arteries was not significantly different between Lean and ZDF rats (33). In our experiments, we found a 2.1-fold reduction of BK channel β1-subunit expression in the aortas of ZDF rats. The discrepancy between these studies may be attributable to the use of vessels from different tissue beds and a difference in methodologies. Although we did not directly determine β1-subunit expression in coronary arteries, the loss of DHS-1-mediated BK channel activation in the coronary artery smooth muscle cells of ZDF rats could be directly explained by a substantial reduction in β1-subunit expression, similar to that in the aorta. We also found that the expression of the BK channel α-subunit in Lean and ZDF rats was the same, and similar to that reported by Burham et al. (33). However, we cannot exclude the possibility that the BK channel α-subunit may be functionally impaired in ZDF rats. The α-subunit is known to be a target of high glucose-mediated oxidative modulation (41), and oxidation of SloC911 was shown to have effects comparable to those of a functional β1-subunit knockout (40).
It requires more than a 50-state model to account completely for all the features of BK channel Ca2+-dependent and voltage-dependent activation (2,49). However, we found that a simplified model (Scheme I, Fig. 5 A) could adequately describe the difference in Ca2+-dependent gating between Lean and ZDF rats, similar to a previous report (50). Here, O2, O3, and O1 represent the longest, intermediate, and shortest open states of the BK channel in Lean rats, whereas O3, O2, and O1 are the longest, intermediate, and shortest open states for ZDF rats. The transition rates from C2 to O2 and from C3 to O3 are faster than those from C1 to O1 in Lean rats. In contrast, the transition from C1 to O1 is fastest in ZDF rats. These channel gating properties suggest that BK channels dwell in longer open states in Lean rats, but have more frequent flickering openings in ZDF rats. According to Scheme I (Fig. 5 A), the total energy barrier for channel activation at any given Ca2+ concentration was 8.6-fold higher in ZDF rats, suggesting that activations of BK channels in ZDF rats are thermodynamically less favorable. In the simplified model, we did not include the transitions between open states, although it was shown that a Ca2+-dependent transition between O1↔O2 and O2↔O3 may exist. However, adding two more transitions between O1↔O2 and O2↔O3 would not change the total energy barrier, and we can easily calculate the apparent Ca2+ dissociation constants of O1↔O2 and O2↔O3 by applying the principle of microscopic reversibility (58) to the resulting closed cycles: O1–C1–C2–O2 and O2–C2–C3–O3. Under equilibrium conditions, the loop equilibrium constants must obey the relationship: KC1–O1 × KO1–O2* = KC1–C2*× KC2–O2 and KC2–O2 × KO2–O3* = KC2–C3 × KC3–O3*, where KC1–O1, KC2–O2, and KC3–O3 are the opening equilibrium constants, and KO1-O2*, KC1–C2*, KO2–O3*, and KC2–C3* are apparent Ca2+ dissociation constants between the respective resting states (or closed states) and open states. With the values of KC1–O1, KC1–C2*, KC2–O2, KC2–C3*, and KC3–O3 from Scheme I (Fig. 5 A; Table 2), KO1–O2* and KO2–O3* can be calculated and shown to be 0.39 μM and 15.52 μM, respectively, in Lean rats, and 0.13 μM and 52.32 μM, respectively, in ZDF rats. The greater values of Ca2+ dissociation constants in ZDF rats indicate that the Ca2+ binding of BK channel open states was less stable in these animals, leading to a shortening of the longer open dwell-time durations, and increasing the frequency of flickering openings. It is worth emphasizing that adding two more transitions among open states in the loop transition model could only increase the probability of a logarithm likelihood ratio test, but it would not change our conclusions in Scheme I (Fig. 5 A).
Subconductance openings were observed in many channels, including BK channels (23,53,56), Shaker channels (59,60), and cystic fibrosis transmembrane conductance regulator channels (61,62). Two mechanisms for subconductance openings have been proposed. The first involves a blockage of channel subunit conduction. It was shown that dendrotoxin I inhibited BK channels and produced subconductance states by a direct interaction with the channel cytoplasmic pore (63). Coexpression of wild-type Slo channels and the tetraethylammonium (TEA)-insensitive Slo mutant channels (mSloY294V) resulted in four levels of subconductance, according to the composition (the wild-type/mSloY294V ratio) of the tetrameric channel, in recordings with TEA in the pipette solution (54,57). The second mechanism involves the individual activation of each tetrameric K+ channel subunit. Using a chimeric Kv2.1 channel construct containing tandem dimers that linked two K+ channel subunits with different voltage-activation thresholds, Chapman et al. found two levels of subconductance when the membrane was depolarized to potentials between the activation thresholds of the subunits (64). The authors concluded that each subunit must change its conformation from a closed state to an open state as the pore undergoes heteromeric conformations in which some subunits are in the open state, but other subunits remain closed, resulting in subconductance openings (64). Under normal conditions, transitions of heteromeric conformational states are too transient to be observed, and the sojourns could only be detected when transitions were slowed (53). The reason that ZDF BK channels undergo more subconductance openings is not understood. Like Kv channels, however, BK channels are tetrameric and allosterically activated by voltage and by intracellular free Ca2+. Hence, BK channel subconductance states may also arise from subunit heteromeric pore conformational changes. Subconductance states would be more frequently discernible as transitions between subunit conformations become slower. This is in agreement with our results, i.e., that there is a higher energy barrier for Ca2+-dependent channel activation in diabetic BK channels. It was suggested that the functional coupling between α-subunits and β1-subunits is a key determinant of BK channel subconductance openings (23). Loss of β1-subunit function would cause impaired channel sensitivity to voltage and intracellular Ca2+ (17–22). Hence, reduced BK channel β1-subunit expression in ZDF rats would favor BK channel subconductance openings. In addition, impaired allosteric transitions between the fully open and fully closed states of BK channels in ZDF rats may contribute to the increase in subconductance openings (55).
To the best of our knowledge, we have presented the first compelling evidence of a physical and functional loss of β1-subunits in native vascular BK channels in ZDF rats. These changes may account for the profound abnormalities in channel gating properties, including Ca2+-dependent activation. Results from this study have important physiological and clinical implications, and may help us better understand the mechanisms that underlie diabetic vascular BK channel dysfunction, and moreover may provide novel insights into the development of approaches for the management of diabetic vasculopathy.
Acknowledgments
We thank Dr. Steven M. Sine for helpful critique of the manuscript. The DHS-1 was kindly provided by Merck & Co.
T.L. is the recipient of a Junior Faculty Award from the American Diabetes Association. This work was supported by grants from the National Institutes of Health (HL-74180 and HL-080118 to H.L.).
Editor: Richard W. Aldrich.
References
- 1.Cox, D. H., J. Cui, and R. W. Aldrich. 1997. Allosteric gating of a large conductance Ca-activated K+ channel. J. Gen. Physiol. 110:257–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Rothberg, B. S., and K. L. Magleby. 2000. Voltage and Ca2+ activation of single large-conductance Ca2+-activated K+ channels described by a two-tiered allosteric gating mechanism. J. Gen. Physiol. 116:75–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Tanaka, Y., K. Koike, and L. Toro. 2004. MaxiK channel roles in blood vessel relaxations induced by endothelium-derived relaxing factors and their molecular mechanisms. J. Smooth Muscle Res. 40:125–153. [DOI] [PubMed] [Google Scholar]
- 4.Knaus, H. G., M. Garcia-Calvo, G. J. Kaczorowski, and M. L. Garcia. 1994. Subunit composition of the high conductance calcium-activated potassium channel from smooth muscle, a representative of the mSlo and slowpoke family of potassium channels. J. Biol. Chem. 269:3921–3924. [PubMed] [Google Scholar]
- 5.Jiang, Z., M. Wallner, P. Meera, and L. Toro. 1999. Human and rodent MaxiK channel β-subunit genes: cloning and characterization. Genomics. 55:57–67. [DOI] [PubMed] [Google Scholar]
- 6.Jiang, Y., A. Pico, M. Cadene, B. T. Chait, and R. MacKinnon. 2001. Structure of the RCK domain from the E. coli K+ channel and demonstration of its presence in the human BK channel. Neuron. 29:593–601. [DOI] [PubMed] [Google Scholar]
- 7.Jiang, Y., A. Lee, J. Chen, M. Cadene, B. T. Chait, and R. MacKinnon. 2002. Crystal structure and mechanism of a calcium-gated potassium channel. Nature. 417:515–522. [DOI] [PubMed] [Google Scholar]
- 8.Schreiber, M., A. Yuan, and L. Salkoff. 1999. Transplantable sites confer calcium sensitivity to BK channels. Nat. Neurosci. 2:416–421. [DOI] [PubMed] [Google Scholar]
- 9.Xia, X. M., X. Zeng, and C. J. Lingle. 2002. Multiple regulatory sites in large-conductance calcium-activated potassium channels. Nature. 418:880–884. [DOI] [PubMed] [Google Scholar]
- 10.Zeng, X. H., X. M. Xia, and C. J. Lingle. 2005. Divalent cation sensitivity of BK channel activation supports the existence of three distinct binding sites. J. Gen. Physiol. 125:273–286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Yusifov, T., N. Savalli, C. S. Gandhi, M. Ottolia, and R. Olcese. 2008. The RCK2 domain of the human BKCa channel is a calcium sensor. Proc. Natl. Acad. Sci. USA. 105:376–381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Zhang, X., C. R. Solaro, and C. J. Lingle. 2001. Allosteric regulation of BK channel gating by Ca2+ and Mg2+ through a nonselective, low affinity divalent cation site. J. Gen. Physiol. 118:607–636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Shi, J., G. Krishnamoorthy, Y. Yang, L. Hu, N. Chaturvedi, D. Harilal, J. Qin, and J. Cui. 2002. Mechanism of magnesium activation of calcium-activated potassium channels. Nature. 418:876–880. [DOI] [PubMed] [Google Scholar]
- 14.Meera, P., M. Wallner, M. Song, and L. Toro. 1997. Large conductance voltage- and calcium-dependent K+ channel, a distinct member of voltage-dependent ion channels with seven N-terminal transmembrane segments (S0–S6), an extracellular N terminus, and an intracellular (S9–S10) C terminus. Proc. Natl. Acad. Sci. USA. 94:14066–14071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Morrow, J. P., S. I. Zakharov, G. Liu, L. Yang, A. J. Sok, and S. O. Marx. 2006. Defining the BK channel domains required for β1-subunit modulation. Proc. Natl. Acad. Sci. USA. 103:5096–5101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Orio, P., Y. Torres, P. Rojas, I. Carvacho, M. L. Garcia, L. Toro, M. A. Valverde, and R. Latorre. 2006. Structural determinants for functional coupling between the β and α subunits in the Ca2+-activated K+ (BK) channel. J. Gen. Physiol. 127:191–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.McManus, O. B., L. M. Helms, L. Pallanck, B. Ganetzky, R. Swanson, and R. J. Leonard. 1995. Functional role of the β subunit of high conductance calcium-activated potassium channels. Neuron. 14:645–650. [DOI] [PubMed] [Google Scholar]
- 18.Meera, P., M. Wallner, Z. Jiang, and L. Toro. 1996. A calcium switch for the functional coupling between α (hslo) and β subunits (Kv,caβ) of maxi K channels. FEBS Lett. 385:127–128. [DOI] [PubMed] [Google Scholar]
- 19.Tanaka, Y., P. Meera, M. Song, H. G. Knaus, and L. Toro. 1997. Molecular constituents of maxi KCa channels in human coronary smooth muscle: predominant α + β subunit complexes. J. Physiol. (Lond.) 502:545–557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Xia, X. M., J. P. Ding, and C. J. Lingle. 1999. Molecular basis for the inactivation of Ca2+- and voltage-dependent BK channels in adrenal chromaffin cells and rat insulinoma tumor cells. J. Neurosci. 19:5255–5264. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Cox, D. H., and R. W. Aldrich. 2000. Role of the β1 subunit in large-conductance Ca2+-activated K+ channel gating energetics. Mechanisms of enhanced Ca2+ sensitivity. J. Gen. Physiol. 116:411–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Giangiacomo, K. M., A. Kamassah, G. Harris, and O. B. McManus. 1998. Mechanism of maxi-K channel activation by dehydrosoyasaponin-I. J. Gen. Physiol. 112:485–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Nimigean, C. M., and K. L. Magleby. 1999. The β subunit increases the Ca2+ sensitivity of large conductance Ca2+-activated potassium channels by retaining the gating in the bursting states. J. Gen. Physiol. 113:425–440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Brenner, R., G. J. Perez, A. D. Bonev, D. M. Eckman, J. C. Kosek, S. W. Wiler, A. J. Patterson, M. T. Nelson, and R. W. Aldrich. 2000. Vasoregulation by the β1 subunit of the calcium-activated potassium channel. Nature. 407:870–876. [DOI] [PubMed] [Google Scholar]
- 25.Pluger, S., J. Faulhaber, M. Furstenau, M. Lohn, R. Waldschutz, M. Gollasch, H. Haller, F. C. Luft, H. Ehmke, and O. Pongs. 2000. Mice with disrupted BK channel β1 subunit gene feature abnormal Ca2+- spark/STOC coupling and elevated blood pressure. Circ. Res. 87:E53–E60. [DOI] [PubMed] [Google Scholar]
- 26.Luscher, T. F., M. A. Creager, J. A. Beckman, and F. Cosentino. 2003. Diabetes and vascular disease: pathophysiology, clinical consequences, and medical therapy: part II. Circulation. 108:1655–1661. [DOI] [PubMed] [Google Scholar]
- 27.De Vriese, A. S., T. J. Verbeuren, J. Van de Voorde, N. H. Lameire, and P. M. Vanhoutte. 2000. Endothelial dysfunction in diabetes. Br. J. Pharmacol. 130:963–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Deinum, J., and N. Chaturvedi. 2002. The renin-angiotensin system and vascular disease in diabetes. Semin. Vasc. Med. 2:149–156. [DOI] [PubMed] [Google Scholar]
- 29.Bubolz, A. H., H. Li, Q. Wu, and Y. Liu. 2005. Enhanced oxidative stress impairs cAMP-mediated dilation by reducing Kv channel function in small coronary arteries of diabetic rats. Am. J. Physiol. Heart Circ. Physiol. 289:H1873–H1880. [DOI] [PubMed] [Google Scholar]
- 30.Kersten, J. R., L. A. Brooks, and K. C. Dellsperger. 1995. Impaired microvascular response to graded coronary occlusion in diabetic and hyperglycemic dogs. Am. J. Physiol. Heart Circ. Physiol. 268:H1667–H1674. [DOI] [PubMed] [Google Scholar]
- 31.Bouchard, J. F., E. C. Dumont, and D. Lamontagne. 1999. Modification of vasodilator response in streptozotocin-induced diabetic rat. Can. J. Physiol. Pharmacol. 77:980–985. [PubMed] [Google Scholar]
- 32.Miura, H., R. E. Wachtel, F. R. Loberiza, Jr., T. Saito, M. Miura, A. C. Nicolosi, and D. D. Gutterman. 2003. Diabetes mellitus impairs vasodilation to hypoxia in human coronary arterioles: reduced activity of ATP-sensitive potassium channels. Circ. Res. 92:151–158. [DOI] [PubMed] [Google Scholar]
- 33.Burnham, M. P., I. T. Johnson, and A. H. Weston. 2006. Reduced Ca2+-dependent activation of large-conductance Ca2+-activated K+ channels from arteries of type 2 diabetic Zucker Diabetic Fatty rats. Am. J. Physiol. Heart Circ. Physiol. 290:H1520–H1527. [DOI] [PubMed] [Google Scholar]
- 34.Zhou, W., X. L. Wang, T. L. Kaduce, A. A. Spector, and H. C. Lee. 2005. Impaired arachidonic acid-mediated dilation of small mesenteric Arteries in Zucker Diabetic Fatty Rats. Am. J. Physiol. Heart. Circ. Physiol. 288:H2210–H2218. [DOI] [PubMed]
- 35.McGahon, M. K., D. P. Dash, A. Arora, N. Wall, J. Dawicki, D. A. Simpson, C. N. Scholfield, J. G. McGeown, and T. M. Curtis. 2007. Diabetes downregulates large-conductance Ca2+-activated potassium β 1 channel subunit in retinal arteriolar smooth muscle. Circ. Res. 100:703–711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Lu, T., X. L. Wang, T. He, W. Zhou, T. L. Kaduce, Z. S. Katusic, A. A. Spector, and H. C. Lee. 2005. Impaired arachidonic acid-mediated activation of large-conductance Ca2+-activated K+ channels in coronary arterial smooth muscle cells in Zucker Diabetic Fatty rats. Diabetes. 54:2155–2163. [DOI] [PubMed] [Google Scholar]
- 37.Dimitropoulou, C., G. Han, A. W. Miller, M. Molero, L. C. Fuchs, R. E. White, and G. O. Carrier. 2002. Potassium (BK(Ca)) currents are reduced in microvascular smooth muscle cells from insulin-resistant rats. Am. J. Physiol. Heart Circ. Physiol. 282:H908–H917. [DOI] [PubMed] [Google Scholar]
- 38.Liu, Y., K. Terata, Q. Chai, H. Li, L. H. Kleinman, and D. D. Gutterman. 2002. Peroxynitrite inhibits Ca2+-activated K+ channel activity in smooth muscle of human coronary arterioles. Circ. Res. 91:1070–1076. [DOI] [PubMed] [Google Scholar]
- 39.Li, H., Q. Chai, D. D. Gutterman, and Y. Liu. 2003. Elevated glucose impairs cAMP-mediated dilation by reducing Kv channel activity in rat small coronary smooth muscle cells. Am. J. Physiol. Heart Circ. Physiol. 285:H1213–H1219. [DOI] [PubMed] [Google Scholar]
- 40.Tang, X. D., M. L. Garcia, S. H. Heinemann, and T. Hoshi. 2004. Reactive oxygen species impair Slo1 BK channel function by altering cysteine-mediated calcium sensing. Nat. Struct. Mol. Biol. 11:171–178. [DOI] [PubMed] [Google Scholar]
- 41.Lu, T., T. He, Z. S. Katusic, and H. C. Lee. 2006. Molecular mechanisms mediating inhibition of human large conductance Ca2+-activated K+ channels by high glucose. Circ. Res. 99:607–616. [DOI] [PubMed] [Google Scholar]
- 42.Lu, T., D. Ye, X. Wang, J. M. Seubert, J. P. Graves, J. A. Bradbury, D. C. Zeldin, and H. C. Lee. 2006. Cardiac and vascular KATP channels in rats are activated by endogenous epoxyeicosatrienoic acids through different mechanisms. J. Physiol. (Lond.) 575:627–644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Lu, T., P. V. Katakam, M. VanRollins, N. L. Weintraub, A. A. Spector, and H. C. Lee. 2001. Dihydroxyeicosatrienoic acids are potent activators of Ca(2+)-activated K(+) channels in isolated rat coronary arterial myocytes. J. Physiol. (Lond.) 534:651–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Colquhoun, D., and F. J. Sigworth. 1995. Fitting and statistical analysis of single channel recording. In Single-Channel Recording, 2nd ed. B. Sakmann, and E. Neher, editors. Plenum Press, New York. 483–587.
- 45.Qin, F., A. Auerbach, and F. Sachs. 1996. Estimating single-channel kinetic parameters from idealized patch-clamp data containing missed events. Biophys. J. 70:264–280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Qin, F., A. Auerbach, and F. Sachs. 2000. Hidden Markov modeling for single channel kinetics with filtering and correlated noise. Biophys. J. 79:1928–1944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Bolton, T. B. 2006. Calcium events in smooth muscles and their interstitial cells: physiological roles of sparks. J. Physiol. (Lond.) 570:5–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Perez, G. J., A. D. Bonev, and M. T. Nelson. 2001. Micromolar Ca2+ from sparks activates Ca2+-sensitive K+ channels in rat cerebral artery smooth muscle. Am. J. Physiol. Cell Physiol. 281:C1769–C1775. [DOI] [PubMed] [Google Scholar]
- 49.Magleby, K. L. 2003. Gating mechanism of BK (Slo1) channels: so near, yet so far. J. Gen. Physiol. 121:81–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.McManus, O. B., and K. L. Magleby. 1991. Accounting for the Ca(2+)-dependent kinetics of single large-conductance Ca2+ -activated K+ channels in rat skeletal muscle. J. Physiol. (Lond.) 443:739–777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Popescu, G., and A. Auerbach. 2003. Modal gating of NMDA receptors and the shape of their synaptic response. Nat. Neurosci. 6:476–483. [DOI] [PubMed] [Google Scholar]
- 52.McManus, O. B., G. H. Harris, K. M. Giangiacomo, P. Feigenbaum, J. P. Reuben, M. E. Addy, J. F. Burka, G. J. Kaczorowski, and M. L. Garcia. 1993. An activator of calcium-dependent potassium channels isolated from a medicinal herb. Biochemistry. 32:6128–6133. [DOI] [PubMed] [Google Scholar]
- 53.Ferguson, W. B., O. B. McManus, and K. L. Magleby. 1993. Opening and closing transitions for BK channels often occur in two steps via sojourns through a brief lifetime subconductance state. Biophys. J. 65:702–714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Niu, X., and K. L. Magleby. 2002. Stepwise contribution of each subunit to the cooperative activation of BK channels by Ca2+. Proc. Natl. Acad. Sci. USA. 99:11441–11446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Moss, G. W., and E. Moczydlowski. 1996. Rectifying conductance substates in a large conductance Ca2+ -activated K+ channel: evidence for a fluctuating barrier mechanism. J. Gen. Physiol. 107:47–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Moss, B. L., S. D. Silberberg, C. M. Nimigean, and K. L. Magleby. 1999. Ca2+-dependent gating mechanisms for dSlo, a large-conductance Ca2+-activated K+ (BK) channel. Biophys. J. 76:3099–3117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Qian, X., X. Niu, and K. L. Magleby. 2006. Intra- and intersubunit cooperativity in activation of BK channels by Ca2+. J. Gen. Physiol. 128:389–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Lauger, P. 1995. Conformational transitions of ionic channels. In Single-Channel Recording, 2nd ed. B. Sakmann and E. Neher, editors. Plenum Press, New York. 651–662.
- 59.Zheng, J., and F. J. Sigworth. 1997. Selectivity changes during activation of mutant Shaker potassium channels. J. Gen. Physiol. 110:101–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Zheng, J., and F. J. Sigworth. 1998. Intermediate conductances during deactivation of heteromultimeric Shaker potassium channels. J. Gen. Physiol. 112:457–474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Tao, T., J. Xie, M. L. Drumm, J. Zhao, P. B. Davis, and J. Ma. 1996. Slow conversions among subconductance states of cystic fibrosis transmembrane conductance regulator chloride channel. Biophys. J. 70:743–753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Harrington, M. A., and R. R. Kopito. 2002. Cysteine residues in the nucleotide binding domains regulate the conductance state of CFTR channels. Biophys. J. 82:1278–1292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Favre, I., and E. Moczydlowski. 1999. Simultaneous binding of basic peptides at intracellular sites on a large conductance Ca2+-activated K+ channel. Equilibrium and kinetic basis of negatively coupled ligand interactions. J. Gen. Physiol. 113:295–320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Chapman, M. L., and A. M. VanDongen. 2005. K channel subconductance levels result from heteromeric pore conformations. J. Gen. Physiol. 126:87–103. [DOI] [PMC free article] [PubMed] [Google Scholar]











