Next Article in Journal
Lipid Remodeling Confers Osmotic Stress Tolerance to Embryogenic Cells during Cryopreservation
Next Article in Special Issue
Editorial to the Special Issue “Electrophysiology”
Previous Article in Journal
sFasL—The Key to a Riddle: Immune Responses in Aging Lung and Disease
Previous Article in Special Issue
In Silico Assessment of Class I Antiarrhythmic Drug Effects on Pitx2-Induced Atrial Fibrillation: Insights from Populations of Electrophysiological Models of Human Atrial Cells and Tissues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Permissive Modulation of Sphingosine-1-Phosphate-Enhanced Intracellular Calcium on BKCa Channel of Chromaffin Cells

1
Krannert Institute of Cardiology and Division of Cardiology, Department of Medicine, Indiana University School of Medicine, Indianapolis, IN 46202, USA
2
Department of Physiology, National Cheng-Kung University Medical College, Tainan 701401, Taiwan
3
Institute of Zoology, National Taiwan University, Taipei 106319, Taiwan
4
Institute of Tissue Engineering & Regenerative Medicine, Helmholtz Zentrum München, 85764 Neuherberg, Germany
5
Division of Pulmonary, Allergy, and Critical Care Medicine, Department of Medicine and Gregory Fleming James Cystic Fibrosis Research Center, University of Alabama at Birmingham, Birmingham, AL 35294, USA
6
Neuroscience Program of Academia Sinica, Academia Sinica, Taipei 115201, Taiwan
7
Institute of Molecular Biology, Academia Sinica, Taipei 115201, Taiwan
8
Institute of Biomedical Sciences, Academia Sinica, Taipei 115201, Taiwan
9
Division of Cardiology, Department of Internal Medicine, Kaohsiung Medical University Hospital, Kaohsiung Medical University, Kaohsiung 807378, Taiwan
10
School of Medicine, College of Medicine, Kaohsiung Medical University, Kaohsiung 807378, Taiwan
11
Graduate Institute of Clinical Medicine, Kaohsiung Medical University, Kaohsiung 807378, Taiwan
12
Department of Pathology, Carver College of Medicine, University of Iowa, Iowa City, IA 52242, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(4), 2175; https://doi.org/10.3390/ijms22042175
Submission received: 15 January 2021 / Revised: 10 February 2021 / Accepted: 18 February 2021 / Published: 22 February 2021
(This article belongs to the Special Issue Electrophysiology)

Abstract

:
Sphingosine-1-phosphate (S1P), is a signaling sphingolipid which acts as a bioactive lipid mediator. We assessed whether S1P had multiplex effects in regulating the large-conductance Ca2+-activated K+ channel (BKCa) in catecholamine-secreting chromaffin cells. Using multiple patch-clamp modes, Ca2+ imaging, and computational modeling, we evaluated the effects of S1P on the Ca2+-activated K+ currents (IK(Ca)) in bovine adrenal chromaffin cells and in a pheochromocytoma cell line (PC12). In outside-out patches, the open probability of BKCa channel was reduced with a mean-closed time increment, but without a conductance change in response to a low-concentration S1P (1 µM). The intracellular Ca2+ concentration (Cai) was elevated in response to a high-dose (10 µM) but not low-dose of S1P. The single-channel activity of BKCa was also enhanced by S1P (10 µM) in the cell-attached recording of chromaffin cells. In the whole-cell voltage-clamp, a low-dose S1P (1 µM) suppressed IK(Ca), whereas a high-dose S1P (10 µM) produced a biphasic response in the amplitude of IK(Ca), i.e., an initial decrease followed by a sustained increase. The S1P-induced IK(Ca) enhancement was abolished by BAPTA. Current-clamp studies showed that S1P (1 µM) increased the action potential (AP) firing. Simulation data revealed that the decreased BKCa conductance leads to increased AP firings in a modeling chromaffin cell. Over a similar dosage range, S1P (1 µM) inhibited IK(Ca) and the permissive role of S1P on the BKCa activity was also effectively observed in the PC12 cell system. The S1P-mediated IK(Ca) stimulation may result from the elevated Cai, whereas the inhibition of BKCa activity by S1P appears to be direct. By the differentiated tailoring BKCa channel function, S1P can modulate stimulus-secretion coupling in chromaffin cells.

1. Introduction

Sphingosine-1-phosphate (S1P), a bioactive sphingolipid, is a blood constituent which elicits numerous biological responses associated with immune cell trafficking [1]. S1P is released by activated platelets [2] and acts as an autocrine or paracrine factor to control many essential cellular functions [3], ranging from a rapid endothelial nitric oxide secretory response to long-term cell growth or death [4,5]. As an extracellular ligand, S1P has been shown to regulate the activity of ionic channels in various cell types. These actions include blocking the Kv1.3 channel in T lymphocytes [6], activating the inward-rectifying K+ current in atrial cardiomyocytes [7], eliciting transient receptor potential channels in vascular smooth muscle cells [8], as well as activating the large-conductance Ca2+-activated K+ channel (BKCa) in endothelial cells [9]. BKCa is sensitive to both increased intracellular Ca2+ (Cai) and/or depolarized membrane potentials (Vm), providing an interplay between the metabolic and electrical state of the cellular activities [10,11]. In conjunction with postganglionic neurons, adrenal chromaffin cells release catecholamine to coordinate the sympathetic response to both environmental and internal stressors. In chromaffin cells, the BKCa activity is critical for controlling exocytosis by altering Vm depolarization and/or Cai mobilization [12,13]. However, whether BKCa is modulated by S1P in these catecholamine-secreting cells remains elusive and the underlying mechanisms are still unknown. In the current study, we performed electrophysiological measurements combined with mathematical modeling to characterize the roles of S1P in BKCa modulation in chromaffin cells. Our data demonstrate that S1P exerts a biphasic effect on BKCa activities through different mechanistic regulations.

2. Results

2.1. S1P Decrease the Single-Channel Currents of BKCa in Cell-Free Mode

We used outside-out patch recording, the cell-free mode, to evaluate the direct effect of S1P on BKCa activities of chromaffin cells bathed in a high-K+ bathing solution containing 1.8 mM CaCl2. S1P (1 µM) significantly reduced the open probability of BKCa channel without changing the single-channel conductance (Figure 1). In contrast, S1P (1 µM) did not have any effects on the BKCa activity in inside-out patches. Figure 1D shows the relationship between the S1P concentration and the percentage inhibition of BKCa open probability. The application of S1P (0.3−30 µM) was found to decrease the probability of channel openings in a concentration-dependent manner. Fitting the dose-response curve with the Hill equation yielded a half-inhibitory concentration of 1.1 µM. In addition, a slope coefficient of 1.2. S1P at a concentration of 30 µM nearly abolished the BKCa activity. The effect of S1P (1 µM) on the kinetics of BKCa gating was further analyzed. In outside-out patches of control, closed time histograms at the level of +60 mV were fitted by a two-exponential curve (Figure 2A). The time constant for the fast and slow components of the closed time histogram were 116 ± 11 and 45 ± 9 ms, respectively (n = 6). When applied to the bath, S1P (1 µM) significantly increased the mean closed times to 167 ± 18 and 63 ± 11 ms (n = 6, p < 0.05). Figure 2B shows the simulated single-channel data generated using transitional rates of BKCa obtained from the experiments. Based on the modeling scheme, the equilibrium dissociation constants in the control and S1P were estimated to be 3.6 ± 0.3 and 3.9 ± 0.3 (n = 5), whereas the equilibrium gating constants in the control and S1P were 0.20 ± 0.07 and 0.033 ± 0.009 (n = 5), respectively. The results suggested that the outer membrane exposure to S1P caused a significant decrease in the gating constant (by about six-folds), but no change in the dissociation constant.

2.2. High-Dose S1P Elevate Intracellular Ca2+ (Cai) Thereby Enhancing BKCa Activities in On-Cell Patches

Next, we measured Cai transients in fura-2-loaded chromaffin cells. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. A representative example of the fura-2 ratio, R (340/380 nm), during the exposure to S1P is shown in Figure 3. S1P (1 µM) did not alter the Cai level, however, a high-dose S1P (10 µM) significantly induced the Cai elevation. An application of high K+ (45 mM) further elevated the Cai intensity. The results indicate that a low-dose of S1P had little or no effect but a high-dose of S1P (10 µM) enhanced Cai. In the cell-attached configuration (on-cell mode), a high-dose S1P (10 µM) significantly increased the opening of BKCa channel (Figure 3C). The open probability of BKCa channel at +60 mV in the absence of S1P was 0.029 ± 0.002 (n = 8). One minute after the high-dose S1P, the open probability was significantly increased to 0.084 ± 0.003 (p < 0.05, n = 12). These channel activities were suppressed by paxilline (Pax, 1 µM), a selective BKCa channel blocker. The single-channel current amplitude and BKCa conductance were unaltered in all the groups, suggesting that BKCa can be induced by Cai at a high depolarization potential (i.e., +60 mV).

2.3. Biphasic Effects of S1P on IK(Ca) of Chromaffin Cells

We evaluated the effects of S1P on IK(Ca) of chromaffin cells by whole-cell voltage-clamp recording. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. Within one minute of exposing the cells to S1P (1 µM), the IK(Ca) was suppressed throughout the entire range of depolarization steps. Representative IK(Ca) traces and averaged current-voltage (I-V) relations for IK(Ca) in the control, during exposure to S1P (1 µM) and washout of S1P are illustrated in Figure 4A, respectively. S1P (1 µM) significantly suppressed IK(Ca), however, this inhibitory effect was readily reversed after removal of S1P. Then, we examined the effect of S1P at a high concentration (10 µM) on IK(Ca). When the cells were exposed to a high-dose S1P (10 µM), a biphasic response in the amplitude of IK(Ca) was observed, i.e., an initial decrease followed by a persistent elevation (Figure 4B). In other words, there was a dual effect of high-dose S1P on IK(Ca) that was elicited by a series of voltage pulses ranging from –30 to +50 mV in 20-mV increments. When the patch pipette was filled with 10 mM BAPTA, a Ca2+ chelator, S1P (10 µM) did not induce an activated IK(Ca). The activation of IK(Ca) by S1P is Cai-dependent which needs to reach the threshold Cai level to be activated.

2.4. Low-Dose S1P Increase Chromaffin Cell Excitability

The effect of S1P on AP firing was investigated by the current-clamp configuration, the cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. The typical effect of S1P (1 µM) on APs is illustrated in Figure 5A. When cells were exposed to S1P, the firing of APs in response to the current stimuli with a duration of 1.6 s was increased. For example, a low-dose S1P significantly increased the firing frequency to 3.1 ± 0.2 Hz (n = 5) from a control of 1.9 ± 0.2 Hz (n = 6). This increase in the firing of AP may result from the IK(Ca) inhibition in chromaffin cells. The AP firing could hardly be generated by 10 µM S1P. Next, we assessed how the decreased conductance of IK(Ca) used to mimic the action of S1P can result in APs changes in modeled chromaffin cells. Figure 5B illustrates the time course of repetitive APs in response to the current injection which was simulated from a simulated cell model. Notably, when the maximal conductance of IK(Ca) was arbitrarily decreased (from 0.64 to 0.32 nS), the firing rate was increased. This result reinforced our observations showing that S1P (1 µM) increased the frequency by two-fold. Concomitant with changes in repetitive firing, AP repolarization was slowed down after a reduction of IK(Ca) conductance which was used to mimic S1P effects.

2.5. S1P Modulate IK(Ca) and BKCa in PC12

The effects of S1P on IK(Ca) and BKCa of pheochromocytoma PC12 cells were further evaluated. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. In whole-cell recording, when the cells were exposed to S1P (1 µM), the IK(Ca) was significantly reduced (Figure 6). Similar to that in chromaffin cells, the BKCa activity was observed in the on-cell patches of PC12 bathed in a high-K+ bathing solution with 1.8 mM CaCl2. In the cell-attached configuration, a high-dose S1P (10 µM) increased the open probability of BKCa channel. This stimulatory effect was not observed in a low-dose S1P (1 µM). The BKCa activities were increased in the presence of squamocin, a Ca2+ ionophore [14], and 10 µM S1P. When Pax (1 µM), a BKCa channel blocker, was applied to the bath, the S1P-stimulated BKCa opening was significantly reversed (Figure 7). Consistent with the recordings in chromaffin cells, S1P (10 µM) can stimulate the BKCa activity expressed in PC12, although 1 µM S1P suppresses IK(Ca).

3. Discussion

The present study demonstrated novel findings that in chromaffin cells, S1P produced biphasic modulations of both inhibitory and stimulatory effects on the BKCa channel depending on different dose levels. The inhibitory effects of S1P (1 µM) on IK(Ca) likely result from BKCa suppression due to the modified lipid component. The activation of BKCa is closely associated with the level of Cai at a high-dose S1P (10 µM). In whole-cell recording, S1P also had a bimodal effect on IK(Ca). Our simulation data predicted that the decreased maximal conductance of the BKCa channel could replicate the experimental data showing that S1P-induced AP firing to stimulate chromaffin cells.

3.1. Endogenous S1P Dosage in Physiology

S1P is abundant in the blood and it can regulate the specific tissues of the whole body. Platelets, which lack the S1P digest enzyme are therefore, rich in S1P and can release S1P upon prothrombotic stimuli, such as thrombin and collagen [15]. The serum S1P concentrations range from 100 nM to 4 μM, and the variation is derived from the different detection methods and by species [16,17,18]. Mouse serum S1P, for example, was measured as 0.84 μM or >4 μM S1P in two different studies, respectively [17,19]. The human serum S1P was reported as 0.3 or 1 μM [16,18]. Moreover, the bioactive concentration of S1P can be changed by its binding lipoproteins, i.e., high- or low-density lipoprotein (HDL or LDL), respectively. The LDL-bound S1P is biologically inactive [20], whereas the HDL-bound S1P is biologically active [21]. The concentration of S1P in blood and plasma constituents are significantly higher than the estimated kDa of S1P to its receptors, which is in the range of 8 to 60 nM [22]. Here, we used the concentrations of 1 and 10 µM to mimic the bioactive status of S1P.

3.2. S1P Modulate BKCa Channel Kinetics in Chromaffin Cells

BKCa in chromaffin cells yielded a single-channel conductance of ~200 pS [13,23]. Under symmetrical K+ conditions, the BKCa activity can be observed in our study similar to previous reports [13,23]. In the outside-out patches, S1P, when applied from the extracellular side on this cell-free mode, was found to be an effective blocker of BKCa. The binding site for S1P on BKCa may be located in the extracellular leaflet. S1P probably associated with its unusual solubility for a lipid due to the polar head-group for the effects of polyamines on BKCa [24]. We demonstrated that S1P can directly decrease the BKCa activity in a dose-dependent fashion with an IC50 value of 1.1 µM. This dosage is very close to normal S1P levels in the serum, which range from 0.4~0.7 µM depending on the species and detect methods differences [4,16,17,18,19]. The physiological dose of S1P had little effect on the mean open time of BKCa, but produced a lengthening in the mean close time. The inhibitory effects of S1P lie primarily on the BKCa kinetics, since there is a difference in the gating constant between, before, and after the S1P application. Our findings indicate that in chromaffin cells, S1P can modify the BKCa gating kinetics leading to a decrease in IK(Ca) for upregulating the cell excitability. Therefore, the BKCa channel can well be an important substrate for S1P in regulating these catecholamine-secreting cells.

3.3. S1P-Induced Cai Elevation Triggers BKCa Activation

In chromaffin cells, extracellular Ca2+ entrance can regulate the BKCa activity [25]. The reasons that S1P stimulates the BKCa activity possibly related to the Cai elevation via either the influx of extracellular Ca2+ into the cytosol through voltage-gated calcium channel (Cav) currents (ICa), or the release of Ca2+ from intracellular stores [26,27]. S1P at a concentration of 1 µM could not induce ICa in chromaffin cells [28]. The S1P-mediated stimulation of BKCa may not result from the activation of Cav. However, whether S1P can activate the transient receptor potential channel remains to be further explored in the neuroendocrine cells, as previously reported [8]. Our data demonstrate that S1P increases the open probability of BKCa in the cell-attached (on-cell) configuration, whereas it produced a reduction in the BKCa activity in the outside-out (cell-free) patches. At the physiological dose level, S1P may affect the cell excitability without any change of Cai. Therefore, a high-dose S1P-triggered BKCa activation can be an inhibition mechanism of Cai elevation for feedback controlling secretion in chromaffin cells.

3.4. Biphasic Effects of S1P on IK(Ca) and the Mechanism of S1P-Mediated Stimulus-Secretion

S1P is metabolized from sphingosine by sphingosine kinases. Sphingosine was shown to have higher potency in triggering endocytosis compared with S1P [29] and the increase of neurotransmitter release has been reported in bovine chromaffin cells [29,30]. S1P regulates exocytosis through distinct mechanisms. S1P can be an extracellular agonist [4,31] and the intracellular S1P contributed a significant extent to the catecholamine release [32]. The extracellular S1P can modulate the rate of exocytosis, while the intracellular S1P may control the fusion pore expansion during exocytosis [32]. In the present study, S1P is complex with biphasic effects of suppressing IK(Ca) at a physiological level, but increase IK(Ca) amplitude at a high-dosage. The stimulatory effect of S1P on IK(Ca) was related to the Cai concentration, whereas inhibitory effects were directly mediated on BKCa, probably extracellularly, although the stimulatory S1P regulation results mainly from the Cai elevation. Based on the experimental observations of S1P-suppressed BKCa increase AP firing, our simulation data mimicked the excitatory of S1P on Vm modulation in a modeled chromaffin cell. Vm depolarization was reported to increase the intracellular S1P production in PC12 cells [33]. Therefore, S1P can act as both an extracellular mediator and intracellular second messenger on the stimulus-secretion coupling by tailoring BKCa. When formed intracellularly in response to extracellular agonists, S1P would be released and modify cell-surface ionic channels [4,5,31,34]. The formation of intracellular S1P and the blockade of BKCa by extracellular S1P may synergistically act on catecholamine secretion of chromaffin cells.

3.5. Possible Mechanisms of S1P-Mediated Regulations on BKCa and Study Limitations

Our observations imply that S1P exerts a bimodal effect on IK(Ca) through different mechanistic regulations of Cai in chromaffin cells. At the tested doses of S1P, BKCa could be modulated by different mechanisms. S1P binds to the S1P receptor (S1PR), the specific cell surface G protein-coupled receptors (GPCRs). The possible regulatory mechanisms of S1P on BKCa may encompass both S1PR-dependent [35] and GPCR-independent [9] modulations. A high-dose S1P may not selectively affect BKCa by the S1PR-dependent regulation, but could possibly regulate other K+ channels, such as the delayed-rectifier K+ current [34], or trigger the Cai elevation as we demonstrated. GTP and G proteins are necessary for the S1P function through GPCR. S1P can activate BKCa in the absence of GTP and in the presence of the G protein inhibitors [9], which suggested that S1P regulates BKCa independently of GPCR. Moreover, the fatty acid-induced modulation of BKCa channel was reported to be lipid-type dependent [36,37,38,39]. Based on these key observations, S1P at a low concentration may work through an S1PR-dependent pathway to regulate BKCa directly, whereas a high S1P may trigger multiple mechanisms, such as GPCR-independent [9] or fatty acid-induced BKCa functional modifications [37]. Nonetheless, the exact mechanism through which S1PR of S1P-mediated BKCa regulation in chromaffin cell deserves further investigation. Five subtypes of S1PRs in the cell surface have been identified, i.e., S1P1 to S1P5 [40]. S1PRs, initially called the endothelial differentiation gene receptors [41], belonged to the members of GPCRs, and it had been reported that S1P1 to S1P3 are expressed in PC12 and/or chromaffin cells [32,42,43,44]. A limitation here is that we did not directly measure the specificity of S1PRs responsible for the S1P-mediated BKCa regulation. This is beyond the scope of this study, however, our findings still provide novel insights into the involvement of S1P in the permissive modulation of Cai on the BKCa channel in the stimulus-secretion coupling of chromaffin cells.
In conclusion, the present study shows that the S1P-mediated permissive IK(Ca) stimulation is indirect and results from an elevated Cai, whereas S1P inhibits the BKCa channel activity directly. S1P can modulate the adrenal chromaffin cell excitability by the differentiated tailoring BKCa channel function.

4. Materials and Methods

4.1. Cell Preparations

Bovine adrenal glands were obtained from a local slaughterhouse and chromaffin cells were prepared by digestion with collagenase (0.5 mg/mL) followed by the density gradient centrifugation, as previously described [28]. These cells were kindly provided by Dr Chien-Yuan Pan (National Taiwan University, Taipei, Taiwan). The chromaffin cells were maintained in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (FBS) in a 5% CO2 at 37 °C. Electrophysiological experiments were carried out between 5 and 10 days after the cells isolated. Pheochromocytoma PC12 cells were obtained from the Bioresources Collection and Research Center (BCRC-60048; Hsinchu, Taiwan) and maintained in the RPMI-1640 medium supplemented with 10% heat-inactivated horse serum and 5% FBS. All the protocols were approved by the National Cheng-Kung University Investigation Committee.

4.2. Drugs and Solutions

D-erythro-sphingosine-1-phosphate (S1P), ionophore, and collagenase were obtained from Sigma-Aldrich (St. Louis, MO, USA). Paxilline was from Biomol Laboratories, Inc. (Plymouth Meeting, PA, USA). BAPTA (bis-(o-aminophenoxy) ethane-N,N,N′,N′-tetraacetic acid tetrakis) and fura-2/AM were obtained from Molecular Probes. Unless otherwise stated, culture media (e.g., DMEM), FBS, horse serum, and trypsin/EDTA were acquired from Thermo Fisher (Waltham, MA, USA), while other chemicals, such as EGTA and HEPES, were of analytical grade. S1P was dissolved in a chloroform:methanol/1:19 solution to a concentration of 1 mM. Then, it was air-dried and dissolved in ethanol to make a stock of 1 mM and stored at −20 °C.
For electrophysiological recordings, the composition of all the extracellular or intracellular solutions (i.e., normal Tyrode solution) used in this work was elaborated in Table 1. In the experiments of the recording IK(Ca) or membrane potential, we used normal Tyrode in the bath and K+-aspartate solution in the pipette. The pipette solution contained low EGTA (0.1 mM). For the single-channel recordings, the high K+ solutions were used in both external and internal milieus. The free Ca2+ concentration of 0.1 µM was calculated assuming a dissociation constant for EGTA and Ca2+ in the pipette solutions.

4.3. Intracellular Ca2+ (Cai) Measurements

Cells were loaded with 5 µM fura-2/AM for 30 min at 25 °C in a normal Tyrode solution. The coverslips on which the cells were grown were mounted in a 1-mL chamber and placed on an inverted fluorescence microscope (DM-IRE2, Leica, Wetzlar, Germany). Cai were monitored with digital imaging using a C-IMAGING system equipped with a Polychrome V high-speed monochromator (Compix Inc, Sewickley, PA, USA). Fura-2/AM was excited sequentially by 340 and 380 nm light delivered from a xenon lamp via a ×20, 1.3 NA UV fluorescence objective (Leica). Fluorescent images were collected at a 510 nm scale every 1.0 s by a Peltier-cooled CCD camera. The intensity of fluorescence ratio, 340/380 nm, from individual cells was measured by the SimplePCI software (Compix).

4.4. Electrophysiological Measurements

Immediately before each experiment, cells were transferred to a chamber positioned on the stage of an inverted microscope (Leica). The glass pipettes were fabricated from Kimax-51 capillaries (Kimble Glass, Vineland, NJ, USA) using a PP-830 puller (Narishige, Japan) and the tips were fire-polished with an MF-83 microforge (Narishige). When filled with a pipette solution, their resistance ranged between 3 and 5 MΩ. Ionic currents were recorded in different patch-clamp configurations (i.e., cell-attached, outside-out, and whole-cell variants), with the use of an RK-400 amplifier (Bio-Logic, Claix, France) controlled by the ClampEx program (Molecular Devices, Sunnyvale, CA, USA). To measure IK(Ca), cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2 at room temperature. Each cell was held at 0 mV to inactivate other non-specific voltage-gated K+ current. The method elicits a family of large, noisy, and outward currents with rectification. When external Ca2+ was removed, the current amplitude was reduced and the residue current was specifically identified as IK(Ca).

4.5. Single-Channel Analyses

Single-channel currents of BKCa channel were analyzed in the pCLAMP 10.4 software (molecular devices). Multi-Gaussian adjustments of the amplitude distributions among channels were used to determine single-channel currents. The functional independence between the channels was verified by comparing the observed stationary probabilities. The probability of channel openings was evaluated using an iterative process to minimize the χ2 calculated with many independent observations. Open or closed lifetime distributions were fitted with a logarithmically scaled bin width. To estimate all the transition rates between states, single-channel data recorded from chromaffin cells were idealized and used to determine single-channel kinetic parameters by means of a maximum likelihood algorithm in the QUB software [45]. The highest log likelihood for the observed BKCa channel activity was obtained with the gating scheme:
C 1 k 2 k 2 C 2 k 1 k 1 O
where O is the open state, and C1 and C2 represent the closed states. On the basis of this linear scheme, single-channel data were modeled, and transition rates were obtained. The gating scheme consists of one open and two closed states. Transition rates between states were derived from maximum likelihood estimations [45]. The kinetic analysis was based on the assumption that channel states in the patch are mutually independent, and the BKCa channel can be described by a finite-state Markovian model.
To calculate the percentage decrease in the BKCa activity resulting from the presence of S1P, the potential was held at +60 mV. When outside-out patches were formed, S1P at different concentrations (0.1−30 µM) was applied to the bath. The concentration-dependent relationship of S1P on the inhibition of BKCa activity in adrenal chromaffin cells was fitted to the Hill equation using a nonlinear regression analysis. That is,
%   d e c r e a s e = E m a x · [ C ] n I C 50 n + [ C ] n
where [C] indicates the S1P concentration; IC50 and n are the concentrations for a 50% inhibition and the Hill coefficient, respectively; and Emax is the S1P-induced maximal inhibition of BKCa channel.

4.6. Statistical Analysis

Averaged results are presented as the mean ± SEM. Paired t-tests were used to compare two different treatments in the same cell. One-way repeated measures ANOVA with post-hoc tests were used to compare the means among three or more different treatments. A two-sided p-value of < 0.05 was considered as statistically significant.

4.7. Computer Simulations

Simulated APs in bovine chromaffin cells were run using the AP firings originally derived from gonadotropin-releasing hormone-secreting neurons [46]. The model consists of a Na+ current, L-type and T-type Ca2+ currents, a delayed rectifier K+ current, and an M-like current. In addition, erg (ether-à-go-go-related-gene) K+ and Ca2+-activated K+ currents were incorporated to this model, given that they are consistently present in chromaffin cells [47,48]. Simulations were carried out using the stochastic algorithm as implemented in the program xpp with the aid of XPPAUT. Source codes were downloaded from http://senselab.med.yale.edu/senselab/modeldb (22 February 2021). Expressions and parameters for ion currents, equilibrium functions, and parameter values are modified to mimic APs in adrenal chromaffin cells observed in this study. The conductance values and reversal potentials, together with other parameters, used to solve the set of all differential equations are listed in Table 2.

Author Contributions

A.Z.W., Y.-C.C., and S.-N.W. contributed to the conception and design of this research; A.Z.W., T.-L.O., H.-C.L., and S.-N.W. performed the experiments and wrote the paper; A.Z.W. and S.-N.W. analyzed the data; A.Z.W., R.-J.S., H.-F.W., H.-C.L., and D.-F.D. reviewed the paper; A.Z.W. and D.-F.D. supported the funding for the project. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by the National Science Council Foundation (NSC 94-2627-M-002-003, NSC-94-2320-B-006-019), the Ministry of Science and Technology (MOST-109-2314-B-037-111-MY3), Kaohsiung Medical University grants (105KMUOR08, KMUH108-8R10, KMUH109-9R11), Academia Sinica grant (AS-CFII-110-101), Taiwan; and the National Institutes of Health (HL145138-02), the Dr Charles Fisch Cardiovascular Research Award (KIC2019FA) endowed by Dr Suzanne Knoebel, a Medtronic-Zipes Endowment, Indiana University School of Medicine (IUSM)-IU Health Strategic Research Initiative, USA.

Acknowledgments

We are indebted to Dr. Peng-Sheng Chen (https://bio.cedars-sinai.org/chenp3/index.html) for his long-term mentorship and the team assistance in Krannert Institute Cardiology, IUSM (A.Z.W.). The authors thank Drs. Yit-Tsong Chen (https://labs.iams.sinica.edu.tw/project/ytchen/pages/professor-yit-tsong-chen) and Chien-Yuan Pan (http://homepage.ntu.edu.tw/~deptlifesci/faculty_ChienYuanPan.html?lang=en) for their help in cell preparation and instrumentation. We also thank Dr. Wen-Bin Liau’s (http://www.mse.ntu.edu.tw/index.php?option=com_zoo&task=item&item_id=108&Itemid=914&lang=en) invaluable opinions for the paper, as well as Dr. Po-Yuan Shih (https://www6.rennes.inrae.fr/igepp_eng/Personnel/S/Shih-Po-Yuan) and Chih-Cheng Yang’s technical help in the pilot study. We are grateful to Dr. Subha Raman (https://medicine.iu.edu/faculty/47775/raman-subha) of IUSM Cardiology for her support in this work.

Conflicts of Interest

The remaining authors declare no conflict of interest.

Abbreviations

ANOVAanalysis of variance
APaction potential
BAPTAbis-(o-aminophenoxy)ethane-N,N,N’,N’-tetraacetic acid tetrakis
BKCalarge-conductance Ca2+-activated K+ channel
Caiintracellular Ca2+ concentration
Cavvoltage-gated Ca2+ channel
EGTAethylene glycol tetraacetic acid
Emaxmaximal inhibition of channel activity
IC50the concentration for a 50% inhibition
ICavoltage-gated Ca2+ currents
IK(Ca)Ca2+-activated K+ currents
Paxpaxilline
PC12pheochromocytoma cell line
S1Psphingosine-1-phosphate
S1PRS1P receptor
Vmmembrane potential

References

  1. Cartier, A.; Hla, T. Sphingosine 1-phosphate: Lipid signaling in pathology and therapy. Science 2019, 366, eaar5551. [Google Scholar] [CrossRef] [PubMed]
  2. Yatomi, Y. Plasma sphingosine 1-phosphate metabolism and analysis. Biochim. Biophys. Acta 2008, 1780, 606–611. [Google Scholar] [CrossRef]
  3. Vito, C.D.; Hadi, L.A.; Navone, S.E.; Marfia, G.; Campanella, R.; Mancuso, M.E.; Riboni, L. Platelet-derived sphingosine-1-phosphate and inflammation: From basic mechanisms to clinical implications. Platelets 2016, 27, 393–401. [Google Scholar] [CrossRef]
  4. Pyne, S.; Pyne, N. Sphingosine 1-phosphate signalling via the endothelial differentiation gene family of G-protein-coupled receptors. Pharmacol. Ther. 2000, 88, 115–131. [Google Scholar] [CrossRef]
  5. Posse de Chaves, E.I. Sphingolipids in apoptosis, survival and regeneration in the nervous system. Biochim. Biophys. Acta 2006, 1758, 1995–2015. [Google Scholar] [CrossRef] [Green Version]
  6. Teisseyre, A.; Michalak, K.; Kuliszkiewicz-Janus, M. The influence of membrane lipid metabolites on lymphocyte potassium channel activity. Cell. Mol. Biol. Lett. 2002, 7, 1095–1109. [Google Scholar]
  7. Ochi, R.; Momose, Y.; Oyama, K.; Giles, W.R. Sphingosine-1-phosphate effects on guinea pig atrial myocytes: Alterations in action potentials and K+ currents. Cardiovasc. Res. 2006, 70, 88–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Xu, S.Z.; Muraki, K.; Zeng, F.; Li, J.; Sukumar, P.; Shah, S.; Dedman, A.M.; Flemming, P.K.; McHugh, D.; Naylor, J.; et al. A sphingosine-1-phosphate-activated calcium channel controlling vascular smooth muscle cell motility. Circ. Res. 2006, 98, 1381–1389. [Google Scholar] [CrossRef]
  9. Kim, M.Y.; Liang, G.H.; Kim, J.A.; Kim, Y.J.; Oh, S.; Suh, S.H. Sphingosine-1-phosphate activates BKCa channels independently of G protein-coupled receptor in human endothelial cells. Am. J. Physiol. Cell Physiol. 2006, 290, C1000–C1008. [Google Scholar] [CrossRef] [PubMed]
  10. Wu, S.N.; Wu, A.Z.; Lin, M.W. Pharmacological roles of the large-conductance calcium-activated potassium channel. Curr. Top. Med. Chem. 2006, 6, 1025–1030. [Google Scholar] [CrossRef]
  11. Gueguinou, M.; Chantome, A.; Fromont, G.; Bougnoux, P.; Vandier, C.; Potier-Cartereau, M. KCa and Ca(2+) channels: The complex thought. Biochim. Biophys. Acta 2014, 1843, 2322–2333. [Google Scholar] [CrossRef] [Green Version]
  12. Gavello, D.; Vandael, D.; Gosso, S.; Carbone, E.; Carabelli, V. Dual action of leptin on rest-firing and stimulated catecholamine release via phosphoinositide 3-kinase-driven BK channel up-regulation in mouse chromaffin cells. J. Physiol. 2015, 593, 4835–4853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lovell, P.V.; James, D.G.; McCobb, D.P. Bovine versus rat adrenal chromaffin cells: Big differences in BK potassium channel properties. J. Neurophysiol. 2000, 83, 3277–3286. [Google Scholar] [CrossRef] [Green Version]
  14. Wu, S.N.; Wang, Y.J.; Lin, M.W. Potent stimulation of large-conductance Ca2+-activated K+ channels by rottlerin, an inhibitor of protein kinase C-delta, in pituitary tumor (GH3) cells and in cortical neuronal (HCN-1A) cells. J. Cell. Physiol. 2007, 210, 655–666. [Google Scholar] [CrossRef]
  15. Yang, L.; Yatomi, Y.; Miura, Y.; Satoh, K.; Ozaki, Y. Metabolism and functional effects of sphingolipids in blood cells. Br. J. Haematol. 1999, 107, 282–293. [Google Scholar] [CrossRef] [PubMed]
  16. Ruwisch, L.; Schafer-Korting, M.; Kleuser, B. An improved high-performance liquid chromatographic method for the determination of sphingosine-1-phosphate in complex biological materials. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2001, 363, 358–363. [Google Scholar] [CrossRef] [PubMed]
  17. Allende, M.L.; Sasaki, T.; Kawai, H.; Olivera, A.; Mi, Y.; van Echten-Deckert, G.; Hajdu, R.; Rosenbach, M.; Keohane, C.A.; Mandala, S.; et al. Mice deficient in sphingosine kinase 1 are rendered lymphopenic by FTY720. J. Biol. Chem. 2004, 279, 52487–52492. [Google Scholar] [CrossRef] [Green Version]
  18. Butter, J.J.; Koopmans, R.P.; Michel, M.C. A rapid and validated HPLC method to quantify sphingosine 1-phosphate in human plasma using solid-phase extraction followed by derivatization with fluorescence detection. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2005, 824, 65–70. [Google Scholar] [CrossRef] [PubMed]
  19. Schwab, S.R.; Pereira, J.P.; Matloubian, M.; Xu, Y.; Huang, Y.; Cyster, J.G. Lymphocyte sequestration through S1P lyase inhibition and disruption of S1P gradients. Science 2005, 309, 1735–1739. [Google Scholar] [CrossRef]
  20. Murata, N.; Sato, K.; Kon, J.; Tomura, H.; Yanagita, M.; Kuwabara, A.; Ui, M.; Okajima, F. Interaction of sphingosine 1-phosphate with plasma components, including lipoproteins, regulates the lipid receptor-mediated actions. Biochem. J. 2000, 352 Pt 3, 809–815. [Google Scholar] [CrossRef]
  21. Okajima, F. Plasma lipoproteins behave as carriers of extracellular sphingosine 1-phosphate: Is this an atherogenic mediator or an anti-atherogenic mediator? Biochim. Biophys. Acta 2002, 1582, 132–137. [Google Scholar] [CrossRef]
  22. Hla, T.; Lee, M.J.; Ancellin, N.; Paik, J.H.; Kluk, M.J. Lysophospholipids--receptor revelations. Science 2001, 294, 1875–1878. [Google Scholar] [CrossRef]
  23. Solaro, C.R.; Prakriya, M.; Ding, J.P.; Lingle, C.J. Inactivating and noninactivating Ca(2+)- and voltage-dependent K+ current in rat adrenal chromaffin cells. J. Neurosci. Off. J. Soc. Neurosci. 1995, 15, 6110–6123. [Google Scholar] [CrossRef]
  24. Weiger, T.M.; Langer, T.; Hermann, A. External action of di- and polyamines on maxi calcium-activated potassium channels: An electrophysiological and molecular modeling study. Biophys. J. 1998, 74 Pt 1, 722–730. [Google Scholar] [CrossRef] [Green Version]
  25. Marty, A. Ca-dependent K channels with large unitary conductance in chromaffin cell membranes. Nature 1981, 291, 497–500. [Google Scholar] [CrossRef]
  26. Ghosh, T.K.; Bian, J.; Gill, D.L. Intracellular calcium release mediated by sphingosine derivatives generated in cells. Science 1990, 248, 1653–1656. [Google Scholar] [CrossRef]
  27. Ghosh, T.K.; Bian, J.; Gill, D.L. Sphingosine 1-phosphate generated in the endoplasmic reticulum membrane activates release of stored calcium. J. Biol. Chem. 1994, 269, 22628–22635. [Google Scholar] [CrossRef]
  28. Pan, C.Y.; Wu, A.Z.; Chen, Y.T. Lysophospholipids regulate excitability and exocytosis in cultured bovine chromaffin cells. J. Neurochem. 2007, 102, 944–956. [Google Scholar] [CrossRef]
  29. Rosa, J.M.; Gandia, L.; Garcia, A.G. Permissive role of sphingosine on calcium-dependent endocytosis in chromaffin cells. Pflug. Arch. Eur. J. Physiol. 2010, 460, 901–914. [Google Scholar] [CrossRef]
  30. Darios, F.; Wasser, C.; Shakirzyanova, A.; Giniatullin, A.; Goodman, K.; Munoz-Bravo, J.L.; Raingo, J.; Jorgacevski, J.; Kreft, M.; Zorec, R.; et al. Sphingosine facilitates SNARE complex assembly and activates synaptic vesicle exocytosis. Neuron 2009, 62, 683–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Spiegel, S.; Milstien, S. Sphingosine 1-phosphate, a key cell signaling molecule. J. Biol. Chem. 2002, 277, 25851–25854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Jiang, Z.J.; Delaney, T.L.; Zanin, M.P.; Haberberger, R.V.; Pitson, S.M.; Huang, J.; Alford, S.; Cologna, S.M.; Keating, D.J.; Gong, L.W. Extracellular and intracellular sphingosine-1-phosphate distinctly regulates exocytosis in chromaffin cells. J. Neurochem. 2019, 149, 729–746. [Google Scholar] [CrossRef] [PubMed]
  33. Alemany, R.; Kleuser, B.; Ruwisch, L.; Danneberg, K.; Lass, H.; Hashemi, R.; Spiegel, S.; Jakobs, K.H.; Meyer zu Heringdorf, D. Depolarisation induces rapid and transient formation of intracellular sphingosine-1-phosphate. FEBS Lett. 2001, 509, 239–244. [Google Scholar] [CrossRef] [Green Version]
  34. Chang, W.T.; Liu, P.Y.; Wu, S.N. Actions of FTY720 (Fingolimod), a Sphingosine-1-Phosphate Receptor Modulator, on Delayed-Rectifier K+ Current and Intermediate-Conductance Ca2+-Activated K+ Channel in Jurkat T-Lymphocytes. Molecules 2020, 25, 4525. [Google Scholar] [CrossRef]
  35. Shen, X.; Zhang, L.; Jiang, L.; Xiong, W.; Tang, Y.; Lin, L.; Yu, T. Alteration of sphingosine-1-phosphate with aging induces contractile dysfunction of colonic smooth muscle cells via Ca2+-activated K+ channel (BKCa) upregulation. Neurogastroenterol. Motil. Off. J. Eur. Gastrointest. Motil. Soc. 2021, e14052. [Google Scholar] [CrossRef]
  36. Chang, H.M.; Reitstetter, R.; Gruener, R. Lipid-ion channel interactions: Increasing phospholipid headgroup size but not ordering acyl chains alters reconstituted channel behavior. J. Membr. Biol. 1995, 145, 13–19. [Google Scholar] [CrossRef]
  37. Clarke, A.L.; Petrou, S.; Walsh, J.V., Jr.; Singer, J.J. Modulation of BK(Ca) channel activity by fatty acids: Structural requirements and mechanism of action. Am. J. Physiol. Cell Physiol. 2002, 283, C1441–C1453. [Google Scholar] [CrossRef]
  38. Denson, D.D.; Wang, X.; Worrell, R.T.; Eaton, D.C. Effects of fatty acids on BK channels in GH(3) cells. Am. J. Physiol. Cell Physiol. 2000, 279, C1211–C1219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Lin, M.W.; Wu, A.Z.; Ting, W.H.; Li, C.L.; Cheng, K.S.; Wu, S.N. Changes in membrane cholesterol of pituitary tumor (GH3) cells regulate the activity of large-conductance Ca2+-activated K+ channels. Chin. J. Physiol. 2006, 49, 1–13. [Google Scholar]
  40. Taha, T.A.; Argraves, K.M.; Obeid, L.M. Sphingosine-1-phosphate receptors: Receptor specificity versus functional redundancy. Biochim. Biophys. Acta 2004, 1682, 48–55. [Google Scholar] [CrossRef]
  41. Lee, M.J.; Van Brocklyn, J.R.; Thangada, S.; Liu, C.H.; Hand, A.R.; Menzeleev, R.; Spiegel, S.; Hla, T. Sphingosine-1-phosphate as a ligand for the G protein-coupled receptor EDG-1. Science 1998, 279, 1552–1555. [Google Scholar] [CrossRef] [PubMed]
  42. Molderings, G.J.; Bonisch, H.; Bruss, M.; Wolf, C.; von Kugelgen, I.; Gothert, M. S1P-receptors in PC12 and transfected HEK293 cells: Molecular targets of hypotensive imidazoline I(1) receptor ligands. Neurochem. Int. 2007, 51, 476–485. [Google Scholar] [CrossRef] [PubMed]
  43. Safarian, F.; Khallaghi, B.; Ahmadiani, A.; Dargahi, L. Activation of S1P(1) receptor regulates PI3K/Akt/FoxO3a pathway in response to oxidative stress in PC12 cells. J. Mol. Neurosci. Mn 2015, 56, 177–187. [Google Scholar] [CrossRef]
  44. Wang, W.; Xiang, P.; Chew, W.S.; Torta, F.; Bandla, A.; Lopez, V.; Seow, W.L.; Lam, B.W.S.; Chang, J.K.; Wong, P.; et al. Activation of sphingosine 1-phosphate receptor 2 attenuates chemotherapy-induced neuropathy. J. Biol. Chem. 2020, 295, 1143–1152. [Google Scholar] [CrossRef]
  45. Qin, F.; Auerbach, A.; Sachs, F. Estimating single-channel kinetic parameters from idealized patch-clamp data containing missed events. Biophys. J. 1996, 70, 264–280. [Google Scholar] [CrossRef] [Green Version]
  46. Van Goor, F.; LeBeau, A.P.; Krsmanovic, L.Z.; Sherman, A.; Catt, K.J.; Stojilkovic, S.S. Amplitude-dependent spike-broadening and enhanced Ca(2+) signaling in GnRH-secreting neurons. Biophys. J. 2000, 79, 1310–1323. [Google Scholar] [CrossRef] [Green Version]
  47. Wu, S.N.; Liu, S.I.; Huang, M.H. Cilostazol, an inhibitor of type 3 phosphodiesterase, stimulates large-conductance, calcium-activated potassium channels in pituitary GH3 cells and pheochromocytoma PC12 cells. Endocrinology 2004, 145, 1175–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Gullo, F.; Ales, E.; Rosati, B.; Lecchi, M.; Masi, A.; Guasti, L.; Cano-Abad, M.F.; Arcangeli, A.; Lopez, M.G.; Wanke, E. ERG K+ channel blockade enhances firing and epinephrine secretion in rat chromaffin cells: The missing link to LQT2-related sudden death? FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2003, 17, 330–332. [Google Scholar] [CrossRef]
Figure 1. Effects of sphingosine-1-phosphate (S1P) on BKCa activities in outside-out patches. (A) Original traces in the control (left) and after the application of S1P (1 µM, right) in the bath. The pipette was filled with a high K+ solution with 0.1 µM Ca2+, the cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2, holding at +60 mV. (B) Amplitude histograms of BKCa obtained in the absence (left) and presence (right) of S1P. Data were fitted by Gaussian distributions with the maximum likelihood method. (C) Averaged I-V relations of BKCa obtained in the absence (⬤) and presence (◯) of S1P. Each point represents the mean ± SEM (n = 6−8). The single-channel conductance between the control and S1P is identical. (D) The dose-response relationship for S1P-induced inhibition of BKCa. The IC50 value, maximally inhibitory percentage, and Hill coefficient were 1.1 µM, 100%, and 1.2, respectively. Each point represents the mean ± SEM (n = 4−9, in five cultures).
Figure 1. Effects of sphingosine-1-phosphate (S1P) on BKCa activities in outside-out patches. (A) Original traces in the control (left) and after the application of S1P (1 µM, right) in the bath. The pipette was filled with a high K+ solution with 0.1 µM Ca2+, the cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2, holding at +60 mV. (B) Amplitude histograms of BKCa obtained in the absence (left) and presence (right) of S1P. Data were fitted by Gaussian distributions with the maximum likelihood method. (C) Averaged I-V relations of BKCa obtained in the absence (⬤) and presence (◯) of S1P. Each point represents the mean ± SEM (n = 6−8). The single-channel conductance between the control and S1P is identical. (D) The dose-response relationship for S1P-induced inhibition of BKCa. The IC50 value, maximally inhibitory percentage, and Hill coefficient were 1.1 µM, 100%, and 1.2, respectively. Each point represents the mean ± SEM (n = 4−9, in five cultures).
Ijms 22 02175 g001
Figure 2. Effects of S1P on BKCa kinetics in outside-out patches (as inset). (A) The effects of S1P on mean closed times of BKCa at +60 mV. The pipette solution containing 0.1 µM Ca2+, the cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2. The closed time histograms in the control (upper) and after the application of S1P (1 µM, lower) are illustrated. The abscissa and ordinate show the logarithm of apparent open or closed times (ms) and the square root of the number of counts, respectively. (B) Simulated single-channel currents in the control (upper) and S1P (lower). Simulation models used to analyze the observed data measured at +60 mV are shown in the upper part of each simulated current. Each horizontal arrow pointing to the left represents the binding of S1P or closing of a channel, and each arrow to the right represents the dissociation of S1P or opening of a channel. The units are µM/ms or /ms. O: Open state; C1 and C2, two closed states.
Figure 2. Effects of S1P on BKCa kinetics in outside-out patches (as inset). (A) The effects of S1P on mean closed times of BKCa at +60 mV. The pipette solution containing 0.1 µM Ca2+, the cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2. The closed time histograms in the control (upper) and after the application of S1P (1 µM, lower) are illustrated. The abscissa and ordinate show the logarithm of apparent open or closed times (ms) and the square root of the number of counts, respectively. (B) Simulated single-channel currents in the control (upper) and S1P (lower). Simulation models used to analyze the observed data measured at +60 mV are shown in the upper part of each simulated current. Each horizontal arrow pointing to the left represents the binding of S1P or closing of a channel, and each arrow to the right represents the dissociation of S1P or opening of a channel. The units are µM/ms or /ms. O: Open state; C1 and C2, two closed states.
Ijms 22 02175 g002
Figure 3. Effects of S1P (10 µM) on Cai and BKCa activation of chromaffin cells. (A) Representative Cai trace of the fluorescence ratio (340/380 nm) from an individual cell loaded with fura-2/AM. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. For Cai, there was no change in the S1P (1 µM) application but a high-dose S1P (10 µM) elevated the Cai transients and was further increased by high K+ (45 mM) subsequently. (B) Summary of S1P effects on Cai. Asterisks indicate significant differences (p < 0.05, * control and ** 10 µM S1P, n = 10, in four cultures). (C) S1P-enhanced BKCa on a cell-attached patch (as inset) measured at +60 mV. The cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2. BKCa events were obtained in the control (left), S1P (10 µM, middle), and after paxilline (Pax, 1 µM, right) combined with S1P which were significantly suppressed.
Figure 3. Effects of S1P (10 µM) on Cai and BKCa activation of chromaffin cells. (A) Representative Cai trace of the fluorescence ratio (340/380 nm) from an individual cell loaded with fura-2/AM. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. For Cai, there was no change in the S1P (1 µM) application but a high-dose S1P (10 µM) elevated the Cai transients and was further increased by high K+ (45 mM) subsequently. (B) Summary of S1P effects on Cai. Asterisks indicate significant differences (p < 0.05, * control and ** 10 µM S1P, n = 10, in four cultures). (C) S1P-enhanced BKCa on a cell-attached patch (as inset) measured at +60 mV. The cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2. BKCa events were obtained in the control (left), S1P (10 µM, middle), and after paxilline (Pax, 1 µM, right) combined with S1P which were significantly suppressed.
Ijms 22 02175 g003
Figure 4. Effects of S1P on IK(Ca) in chromaffin cells. (A) Superimposed current traces (left panel) obtained in the control, exposure to S1P (1 µM), and washout. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. Averaged I-V relationships of IK(Ca) measured at the end of each voltage pulse (right panel) in the control (⬤), exposure to 1 µM S1P (◯), and washout (⏹). Each point represents the mean ± SEM (n = 6−10, in four cultures). (B) Biphasic effects of high-dose S1P (10 µM) on IK(Ca). Superimposed current traces obtained when the recording pipette was filled with ethylene glycol tetraacetic acid (EGTA, 0.1 mM). The uppermost part is the control. Time courses after the S1P (10 µM) application are at 1 (a) and 3 min (b). The pulse protocol is shown as an inset part of each related graph.
Figure 4. Effects of S1P on IK(Ca) in chromaffin cells. (A) Superimposed current traces (left panel) obtained in the control, exposure to S1P (1 µM), and washout. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. Averaged I-V relationships of IK(Ca) measured at the end of each voltage pulse (right panel) in the control (⬤), exposure to 1 µM S1P (◯), and washout (⏹). Each point represents the mean ± SEM (n = 6−10, in four cultures). (B) Biphasic effects of high-dose S1P (10 µM) on IK(Ca). Superimposed current traces obtained when the recording pipette was filled with ethylene glycol tetraacetic acid (EGTA, 0.1 mM). The uppermost part is the control. Time courses after the S1P (10 µM) application are at 1 (a) and 3 min (b). The pulse protocol is shown as an inset part of each related graph.
Ijms 22 02175 g004
Figure 5. Effects of S1P on action potentials (APs) and simulated AP of a modeling chromaffin cell. (A) Current-clamp conditions were performed with the current injection of 2 nA applied to the cell with a duration of 1.6 s. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. Potential traces were recorded in the control and exposure to S1P (1 µM). (B) The simulated APs were generated from a cell model which mimics the electrical behavior of chromaffin cell. A current injection at the 4.6 nA with 1.4 s duration was applied. S1P (1 µM) was mimicked IK(Ca) by the maximal conductance (gK(Ca)) from 0.64 (a) to 0.32 nS (b).
Figure 5. Effects of S1P on action potentials (APs) and simulated AP of a modeling chromaffin cell. (A) Current-clamp conditions were performed with the current injection of 2 nA applied to the cell with a duration of 1.6 s. The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2. Potential traces were recorded in the control and exposure to S1P (1 µM). (B) The simulated APs were generated from a cell model which mimics the electrical behavior of chromaffin cell. A current injection at the 4.6 nA with 1.4 s duration was applied. S1P (1 µM) was mimicked IK(Ca) by the maximal conductance (gK(Ca)) from 0.64 (a) to 0.32 nS (b).
Ijms 22 02175 g005
Figure 6. Inhibitory effect of S1P on IK(Ca) in PC12 cells. (A) Superimposed current traces obtained in the control and presence of S1P (1 µM). The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. (B) Averaged I-V relations of current amplitudes measured at the end of voltage pulses in the absence (◯) and presence (⬤) of 1 µM S1P. Each point represents the mean ± SEM (n = 6, in three cultures).
Figure 6. Inhibitory effect of S1P on IK(Ca) in PC12 cells. (A) Superimposed current traces obtained in the control and presence of S1P (1 µM). The cells were bathed in a normal Tyrode solution with 1.8 mM CaCl2, and the pipette solution containing 0.1 µM Ca2+. (B) Averaged I-V relations of current amplitudes measured at the end of voltage pulses in the absence (◯) and presence (⬤) of 1 µM S1P. Each point represents the mean ± SEM (n = 6, in three cultures).
Ijms 22 02175 g006
Figure 7. S1P-induced BKCa activation in the cell-attached configuration of PC12 cells. (A) The cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2, holding at +60 mV. The original current traces were obtained in the control and 3 min after the S1P (10 µM) application. (B) Summary of the effects of squamocin (Squa, 10 µM), S1P, and S1P combined with paxilline (Pax, 1 µM) on the BKCa open probability. The BKCa activity in the control was considered to be 1.0 and the relative open probability by each agent was plotted. Asterisks indicate significant differences (p < 0.05, * control, and ** S1P alone group, n = 5−9, in three cultures).
Figure 7. S1P-induced BKCa activation in the cell-attached configuration of PC12 cells. (A) The cells were bathed in a high-K+ bathing solution with 1.8 mM CaCl2, holding at +60 mV. The original current traces were obtained in the control and 3 min after the S1P (10 µM) application. (B) Summary of the effects of squamocin (Squa, 10 µM), S1P, and S1P combined with paxilline (Pax, 1 µM) on the BKCa open probability. The BKCa activity in the control was considered to be 1.0 and the relative open probability by each agent was plotted. Asterisks indicate significant differences (p < 0.05, * control, and ** S1P alone group, n = 5−9, in three cultures).
Ijms 22 02175 g007
Table 1. The composition of extracellular and intracellular solutions used in this study.
Table 1. The composition of extracellular and intracellular solutions used in this study.
Solution’s NameMilieuComposition (in mM)
Normal Tyrode solutionExtracellularNaCl 136.5, KCl 5.4, CaCl2 1.8, MgCl2 0.53, glucose 5.5, and HEPES 5.5, adjusted with NaOH to pH 7.4
High K+-bathing solutionExtracellularKCl 145, CaCl2 1.8, MgCl2 0.53, and HEPES 5, adjusted with KOH to pH 7.4
K+-aspartate solutionIntracellularK+-aspartate 130, KCl 20, KH2PO4 1, MgCl2 1, EGTA 0.1, ATP 3, GTP 0.1, and HEPES 5, adjusted with KOH to pH 7.2
High K+-pipette solutionIntracellularKCl 145, MgCl2 2, EGTA 0.1, and HEPES 5, adjusted with KOH to pH 7.2
Table 2. Default parameter values used for the modeling of adrenal chromaffin cells.
Table 2. Default parameter values used for the modeling of adrenal chromaffin cells.
SymbolDescriptionValue
CmMembrane capacitance14 pF
gNaNa+ current conductance65 nS
gCa,LL-type Ca2+ current conductance0.98 nS
gCa,TT-type Ca2+ current conductance0.94 nS
gK(DR)K+ current conductance97 nS
gK(Ca)Ca2+-activated K+ current conductance0.64 nS
gK(M)M-type K+ current conductance0.43 nS
gK(erg)erg K+ current conductance0.95 nS
VNaNa+ reversal potential+60 mV
VCaCa2+ reversal potential+100 mV
VKK+ reversal potential−80 mV
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, A.Z.; Ohn, T.-L.; Shei, R.-J.; Wu, H.-F.; Chen, Y.-C.; Lee, H.-C.; Dai, D.-F.; Wu, S.-N. Permissive Modulation of Sphingosine-1-Phosphate-Enhanced Intracellular Calcium on BKCa Channel of Chromaffin Cells. Int. J. Mol. Sci. 2021, 22, 2175. https://doi.org/10.3390/ijms22042175

AMA Style

Wu AZ, Ohn T-L, Shei R-J, Wu H-F, Chen Y-C, Lee H-C, Dai D-F, Wu S-N. Permissive Modulation of Sphingosine-1-Phosphate-Enhanced Intracellular Calcium on BKCa Channel of Chromaffin Cells. International Journal of Molecular Sciences. 2021; 22(4):2175. https://doi.org/10.3390/ijms22042175

Chicago/Turabian Style

Wu, Adonis Z., Tzu-Lun Ohn, Ren-Jay Shei, Huei-Fang Wu, Yong-Cyuan Chen, Hsiang-Chun Lee, Dao-Fu Dai, and Sheng-Nan Wu. 2021. "Permissive Modulation of Sphingosine-1-Phosphate-Enhanced Intracellular Calcium on BKCa Channel of Chromaffin Cells" International Journal of Molecular Sciences 22, no. 4: 2175. https://doi.org/10.3390/ijms22042175

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

Article Metrics

Back to TopTop