Abstract
Many neurons including vasopressin (VP) magnocellular neurosecretory cells (MNCs) of the hypothalamic supraoptic nucleus (SON) generate afterhyperpolarizations (AHPs) during spiking to slow firing, a phenomenon known as spike frequency adaptation. The AHP is underlain by Ca2+-activated K+ currents, and while slow component (sAHP) features are well described, its mechanism remains poorly understood. Previous work demonstrated that Ca2+ influx through N-type Ca2+ channels is a primary source of sAHP activation in SON oxytocin neurons, but no obvious channel coupling was described for VP neurons. Given this, we tested the possibility of an intracellular source of sAHP activation, namely, the Ca2+-handling organelles endoplasmic reticulum (ER) and mitochondria in male and female Wistar rats. We demonstrate that ER Ca2+ depletion greatly inhibits sAHPs without a corresponding decrease in Ca2+ signal. Caffeine sensitized AHP activation by Ca2+. In contrast to ER, disabling mitochondria with CCCP or blocking mitochondria Ca2+ uniporters (MCUs) enhanced sAHP amplitude and duration, implicating mitochondria as a vital buffer for sAHP-activating Ca2+. Block of mitochondria Na+-dependent Ca2+ release via triphenylphosphonium (TPP+) failed to affect sAHPs, indicating that mitochondria Ca2+ does not contribute to sAHP activation. Together, our results suggests that ER Ca2+-induced Ca2+ release activates sAHPs and mitochondria shape the spatiotemporal trajectory of the sAHP via Ca2+ buffering in VP neurons. Overall, this implicates organelle Ca2+, and specifically ER–mitochondria-associated membrane contacts, as an important site of Ca2+ microdomain activity that regulates sAHP signaling pathways. Thus, this site plays a major role in influencing VP firing activity and systemic hormonal release.
Keywords: afterhyperpolarization, ER, hypothalamus, ion channel, mitochondria, vasopressin
Significance Statement
The slow afterhyperpolarization (sAHP) is mediated by a Ca2+-dependent K+ current. Despite its critical role in regulating neuronal spiking, the Ca2+-dependent mechanisms leading to its activation and spatiotemporal shape remains poorly understood. Here we show that in vasopressin (VP) neurons, dynamic interactions in Ca2+ handling between endoplasmic reticulum (ER) and mitochondria play a significant role in sAHP initiation (via ER Ca2+ release) and its spatiotemporal waveform (via mitochondrial Ca2+ uptake). Our results suggest that contact sites between ER and mitochondria represent Ca2+ microdomains critically involved in initiating the first steps of sAHP generation in VP neurons. Given that changes in the sAHP have been linked to abnormal firing activity in various diseases, our results have both wide-range physiological and pathological implications.
Introduction
Magnocellular neurosecretory cells (MNCs) display different Ca2+-dependent hyperpolarizing afterpotentials that lower the intrinsic excitability of these neurons at different orders of magnitude of activation and duration such as the hyperpolarized afterpotential (HAP) and the afterhyperpolarization (AHP; Roper et al., 2003). The HAP is a hyperpolarizing overshoot of membrane potential lasting ∼25–125 ms and is sustained by an active conductance (Andrew and Dudek, 1984a; Bourque et al., 1985). In contrast, the AHP is an intrinsic excitability mechanism underlain by multiple Ca2+-dependent K+ currents that shape firing frequency on a longer timescale, shaping the patterning of spikes in many types of neurons (Andrade et al., 2012). It is usually described as having three components, underlain by AHP currents (IAHP). The first is the fast AHP (fAHP) mediated by big conductance Ca2+-dependent K+ (BK) channels, which activate after single spikes and deactivate almost as quickly (Dopico et al., 1999). The second is the medium AHP (mAHP) which begins to appear after 3–5 spikes and has a decay time constant of ∼250 ms in VP neurons (Teruyama and Armstrong, 2005). The mAHP current (ImAHP) is underlain by small conductance Ca2+-activated K+ (SK) channels and are sensitive to the bee venom apamin (Bourque and Brown, 1987; Schwindt et al., 1988a,b; Lorenzon and Foehring, 1992). The third component is the apamin-insensitive slow AHP (sAHP) underlain by a sAHP current (IsAHP; Ghamari-Langroudi and Bourque, 2004; Villalobos et al., 2004; Armstrong et al., 2019). The sAHP duration lasts from seconds to tens of seconds and is Ca2+ dependent, is voltage independent, is insensitive to many K+ channel blockers, appears only after longer trains of spikes, and is strongly attenuated by neurotransmitters (Ghamari-Langroudi and Bourque, 2004; Andrade et al., 2012). The mechanisms for sAHP generation differ between neuronal cell types, and the ion channel(s) responsible remains elusive (Andrade et al., 2012). The sAHP is responsible for spike frequency adaptation and is impaired in many neurological diseases and conditions where neuronal firing is altered such as epilepsy and Alzheimer's disease (Moyer et al., 1992; McCormick and Contreras, 2001; Soh and Tzingounis, 2010; Kaczorowski et al., 2011). Thus, understanding sAHP molecular and ionic mechanisms is of critical physiological and clinical relevance.
In oxytocin (OT) and vasopressin (VP) MNCs of the supraoptic nucleus (SON), the sAHP activates during prolonged spike trains and plays a critical role in terminating the respective burst and phasic firing patterns that these neurons exhibit (Greffrath et al., 1998; Ghamari-Langroudi and Bourque, 2004; Teruyama et al., 2008). These patterns optimize neuropeptide release, minimizing fatigue by giving the neuron a cyclical refractory period (Dutton and Dyball, 1979; Bicknell and Leng, 1981; Bicknell, 1988). In VP neurons, the pattern is asynchronous between neurons, resulting in a smooth, sustained tone of VP release from the pituitary (Poulain and Wakerley, 1982; Bourque et al., 1998). As axonal hormone release tightly correlates with MNC activity, this mechanism consequentially shapes the pattern, frequency, and volume of hormones into the bloodstream from MNC axons terminating in the posterior pituitary (Bicknell and Leng, 1981; Cazalis et al., 1985; Armstrong et al., 1986; Gainer et al., 1986).
A rise in intracellular cytosolic Ca2+ concentration ([Ca2+]c) is the first step in activating AHPs (Bourque et al., 1985; Lancaster and Adams, 1986; Ghamari-Langroudi and Bourque, 2004). While mAHPs are generated by direct Ca2+-gated SK activation (Adelman et al., 2012), sAHP mechanisms are slower and likely reflect recruitment of a signal transduction pathway (Vogalis et al., 2003; Villalobos et al., 2004; Andrade et al., 2012; Kim et al., 2016). The specific class of voltage-gated Ca2+ channels (VGCCs) contributing to AHP activation varies widely between neuronal cell types (Andrade et al., 2012). Previous work has demonstrated that N-type account for the majority of mAHP activation in OT and VP neurons (Kirchner et al., 2018). However, a clear cell-type discrepancy exists regarding activation of the sAHP between OT and VP neurons. While N-type VGCCs are the primary source of sAHP Ca2+ in OT neurons (Kirchner et al., 2018), the mechanism in VP neurons remains more complicated, as pharmacological block of any individual high voltage activated (HVA) VGCC failed to account for a majority of sAHP activation (Kirchner et al., 2018).
Three potential sources of activating Ca2+ remain untested. Either the sAHP reflects activation by low voltage activated T-type channels, a proportional rise in global [Ca2+]c, or activation by an internal source of Ca2+ such as the endoplasmic reticulum (ER) or mitochondria. Given that the sAHP activates only after long spike trains, slow kinetics renders T-type Ca2+ an unlikely mechanism, as this channel has fast, transient biophysical properties (Erickson et al., 1993; Sato-Numata et al., 2020). Global [Ca2+]c also seems unlikely given that previous work showed VGCC toxins significantly blocked (33–60%) [Ca2+]c transients yet demonstrated no proportional block of the sAHP (Kirchner et al., 2018). The R-type blocker SNX-482 showed the greatest reduction, and the N-type blocker had only a minor yet statistically significant effect. Both N-type and R-Type blockers demonstrated an inconsistent corresponding reduction in Ca2+ amplitude. Thus, it appears that the sAHP time course does not simply reflect the trajectory of [Ca2+]c. Studies have demonstrated that Ca2+ release from the ER activates the sAHP in other cell types (Sah and McLachlan, 1991; Torres et al., 1996; Shah and Haylett, 2000; Vogalis et al., 2001). Additionally, mitochondria can not only release Ca2+ to the cytosol (Takeuchi et al., 2015), but it also potently buffers [Ca2+]c via the activity of the mitochondrial Ca2+ uniporter (Rizzuto et al., 2009, 2012), including in VP MNCs (Komori et al., 2010; Dayanithi et al., 2012). Furthermore, mitochondrial membranes form complexes with ER membranes (Rizzuto et al., 2009), including with RyRs (Szalai et al., 2000; Hajnóczky et al., 2002), and ER–mitochondria dynamic interactions have been shown to play an important role in regulating neuronal Ca2+ homeostasis (Rizzuto et al., 2009; Giorgi et al., 2015; Karagas and Venkatachalam, 2019). Given these collective findings, we tested here the role of these Ca2+-handling organelles in shaping sAHPs and firing properties in VP neurons.
Materials and Methods
Animals
All experiments were ethically approved by the Georgia State University Institutional Animal Care and Use Committee (IACUC) and performed in accordance with National Institutes of Health's Guide for the Care and Use of Laboratory Animals. All experiments were performed in 30 male and 24 female young adult Wistar rats in the late peripubertal to adolescent developmental stage (150–270 g) with transgenic expression of eGFP tagged to VP (Ueta et al., 2005), which were pooled for analysis. These rats received ad libitum food and water and housed on a standard 12 h light/dark cycle.
Ex vivo slice preparation
On the day of the experiment, rats were anesthetized with pentobarbital (50 mg kg−1, i.p.) and then transcardially perfused with 30 ml of ice-cold sucrose artificial cerebrospinal fluid (aCSF) solution. This sucrose aCSF solution contained the following (in mM): 200 sucrose, 2.5 KCl, 1 MgSO4, 26 NaHCO3, 1.25 NaH2PO4, 20 d-glucose, 0.4 ascorbic acid, and 2.0 CaCl2, pH 7.2, 300–305 mOsmol L−1. The animal was then rapidly decapitated, and the brain was subsequently removed, mounted in the chamber of a vibratome (Leica VT1200S, Leica Microsystems) with superglue, and submerged into sucrose aCSF and bubbled constantly with 95% O2/5% CO2. Slices were cut at 240 µm thickness, bisected at the midline to separate the bilateral SONs into their own separate slice, and placed in a holding chamber containing standard aCSF (described below) bubbled with 95% O2/5% CO2. Slices rested in a water bath at 32°C for 20 min before transfer to room temperature for a minimum of 40 min before any recording.
Whole-cell patch clamp
External recording solution (aCSF) was the following (in mM): 119 NaCl, 2.5 KCl, 1 MgSO4, 26 NaHCO3, 1.25 NaH2PO4, 20 d-glucose, 0.4 ascorbic acid, and 2.0 CaCl2, 2 Na+ pyruvate with 5 mM CsCl added to block the slow depolarizing afterpotential, which overlaps in time course with the AHP (Ghamari-Langroudi and Bourque, 1998; Armstrong et al., 2010). Slices were placed in the chamber of an upright microscope and perfused constantly by aCSF warmed to 32°C at a flow rate of 2–3 ml min−1. All AHP recordings were performed in the presence of synaptic blockers 10 µM 6,7-dinitroquinoxaline-2,3-dione (DNQX), 40 µM 2R)-amino-5-phosphonovaleric acid (AP5), and 100 µM picrotoxin to block fast synaptic currents (Tocris Bioscience). Current-clamp and voltage-clamp recordings were acquired using a MultiClamp 700B and digitized with a Digidata 1550b (Molecular Devices). Electrodes were pulled using a Flaming Brown horizontal puller (Sutter Instruments) from borosilicate capillaries (2.5–5 MΩ) and filled with internal solution (in mM): 135 KMeSO4, 8 KCl, 10 HEPES, 2 Mg-ATP, 0.3 Na-GTP, 6 phosphocreatine, as well as 0.05 fluo-5F for Ca2+ imaging, 7.2 pH [285–295 mosmol (kg H2O)−1]. Once we achieved whole-cell configuration, we allowed the cell to rest for 5 min to ensure stability of the patch and equilibration between internal solution and cytosol. We took baseline and experimental recordings in the presence of CsCl and synaptic blockers. In current clamp, we adjusted holding current to bring the cell to a resting membrane potential of −55 mV. We evoked AHPs using two different categories of protocol. We either used a 2.5 s duration depolarizing current step (0–150 pA) or a pulse train consisting of 30 spikes at 20 Hz with a 5 ms pulse width. Data was acquired at 10 kHz. In voltage clamp, we evaluated IAHP by stepping from −55 to +10 mV at 20 Hz for 30 pulses (5 ms pulse width). Data was acquired at 2 kHz. Cells that displayed a shift in series resistance that exceeded 20 MΩ or a 25% change from the start of the recording were discarded. Experimental reagents include 3 µM thapsigargin (TG; Tocris Bioscience), 10 mM caffeine (Sigma-Aldrich), 1 µM carbonyl cyanide m-chlorophenyl hydrazone (CCCP; Sigma-Aldrich), 20 µM Ru360 (Ru360; EMD Millipore), and 20 µM triphenylphosphonium (TPP+; Sigma-Aldrich).
Calcium imaging
Ca2+ imaging was performed simultaneously with whole-cell patch-clamp using an Andor Dragonfly 200 spinning disk confocal microscope system (Andor Technology). Ca2+ was evaluated via fluo-5F delivered through the patch pipette (0.05 mM). Calcium fluorescence was obtained by exciting fluo-5F (or fluo-4 in uncaging experiments) at 488 nm. We measured fluorescence changes at an emission wavelength of 520 nm.
Photolytic calcium uncaging
Ca2+ uncaging experiments were performed simultaneously with patch clamp and Ca2+ imaging. The pipette internal solution was the same as described above with a few key differences. First, 2 mM DM-Nitrophen Ca2+ caging compound was added along with 0.8 mM CaCl2 (40% occupancy). Second, 2 mM Na-ATP replaced 2 mM Mg-ATP as Mg2+ competes strongly with Ca2+ for DM-Nitrophen binding (Faas et al., 2005). Third, fluo-4 (Kd = 0.35 µM) was used as the Ca2+ indicator instead of fluo-5F (Kd = 2.3 µM), as fluo-5F is unable to compete with DM-Nitrophen with its low Ca2+ binding affinity. After achieving whole-cell configuration, we allowed internal solution to reach equilibrium with the inside of the cell (5 min minimum) before proceeding with experiments. We photolytically activated the DM-Nitrophen with a 50 ms pulse of 405 nm light delivered to the entire field. There was a 2 min minimum latency between trials to ensure membrane potential and [Ca2+]c returned to baseline.
Experimental design and statistical analyses
In MNCs the fast AHP (fAHP) is a rapid event lasting <15 ms (Dopico et al., 1999) and was not evaluated herein. We measured AHP amplitude at two time points after a given stimulus. Since the ImAHP decay tau is ∼500 ms in MCNs (Teruyama and Armstrong, 2005), we evaluated AHP (current clamp) and IAHP (voltage clamp) amplitudes at 0.1 s (mAHP + sAHP) and 1 s (sAHP) from the end of the pulse depolarization or pulse train (Teruyama and Armstrong, 2005). This is to measure potential effects of the early phase of the sAHP (0.1 s after pulse termination) while also capturing a phase known to contain no mAHP contribution (1 s after pulse termination). Duration of the AHP in current clamp is the time from 0.1 s after the pulse to when the potential reaches baseline potential. Area of the AHP and IAHP is the integrated area under the curve. Neuronal gain was calculated by fitting a linear regression to spike frequency data (f-I slope). Data was analyzed in ClampFit 11 (Molecular Devices) and statistical analysis was performed in GraphPad Prism (Dotmatics).
Fluorescent Ca2+ imaging data was background subtracted and measured at the soma excluding the nucleus. All Ca2+ imaging data are reported as percentage relative change in Ca2+ indicator fluorescence divided by baseline fluorescence calculated from a mean of points preceding the stimulus (%ΔF / F). Peak Ca2+ is the highest Ca2+ response point. Ca2+ amplitude is the height of the Ca2+ peak relative to baseline. Ca2+ area is the integrated area under the curve. Raw data was extracted using FIJI (NIH), conversion to %ΔF/F was done via MatLab script (MathWorks) and measurement of waveforms was done in either Igor Pro 9.0 (WaveMetrics) or ClampFit 11. Statistical analysis was performed in GraphPad Prism.
All data was tested for normality using a D'Agostino–Pearson omnibus test and homogeneity of variance. If the data met these conditions, an appropriate parametric test was performed. Otherwise, an equivalent nonparametric test was performed instead. Full statistical test and designs for each experiment can be found in the corresponding figure legend. n values represent the number of cells. No more than one cell per slice was evaluated. Histograms and plots were generated in GraphPad Prism. Example traces were generated by Igor Pro 9. Figures were created in Illustrator (Adobe).
Results
Depletion of ER Ca2+ stores abolishes sAHPs in VP neurons
Previous work studying HVA VGCC type contributions to AHPs in SON VP neurons demonstrated that while N-type channels coupled to the mAHP, no one channel gated the sAHP (Kirchner et al., 2018). Here we tested the contribution of ER Ca2+ as a source to sAHP activation, given its major role in amplifying Ca2+ signals via CICR (Verkhratsky and Shmigol, 1996). We incubated slices in thapsigargin (TG, 3 µM) for 1 h prior to recording to deplete ER Ca2+ stores. We generated AHPs with a pulse of 30 spikes at 20 Hz to control for spike count and Ca2+ influx. TG depletion significantly inhibited evoked sAHP amplitude, area, and duration in VP neurons, while leaving [Ca2+]c peak and area unaffected in both current clamp (Fig. 1A,B) and voltage clamp (Fig. 1C,D). We considered whether biological sex might influence the TG effect on sAHPs, and thus we analyzed the data controlling for sex. TG blocked the sAHP in VP neurons in both males and females (Extended Data Fig. 1-1). Therefore, data collected from both sexes was pooled in each experiment throughout the rest of this study. sAHP recordings from putative OT (eGFP negative) neurons were unaffected by TG (Fig. 1), confirming the previous report that OT sAHPs are activated primarily by Ca2+ influx through N-type VGCCs (Kirchner et al., 2018). Furthermore, TG inhibition of sAHPs translated to meaningful changes in firing discharge, increasing firing rate in VP neurons. TG significantly increased evoked frequency and neuronal gain (f-I slope) in response to single depolarizing steps (Fig. 1E,F). These results suggest that ER Ca2+ gates the activation of sAHPs in VP neurons.
Figure 1.
Depletion of ER Ca2+ stores inhibits sAHPs in VP neurons. See Extended Data Figure 1-1 for an analysis of sex differences. A, Representative AHPs (arrows) recorded in current clamp (top) and corresponding Ca2+ responses (bottom) from four different groups: VP neurons under control conditions, VP neurons after 1 h (TG) incubation, OT neurons under control conditions, and OT neurons after 1 h of TG incubation. Note the inhibited AHP in the TG-incubated VP neuron (asterisk) but lack of effect in the OT neurons. B, Summary data of TG's effects on current-clamp AHPs evoked by a 20-spike 20 Hz depolarizing pulse train from −55 mV. Control (VP) n = 20, TG (VP) n = 40, control (OT) n = 14, TG (OT) n = 12. Statistics performed with a two-way ANOVA and Šídák's multiple comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. 0.1 s amplitude (F = 8.49; p < 0.01), 1 s amplitude (F = 28.0; p < 0.0001), duration (F = 34.0; p < 0.0001), area (F = 30.3; p < 0.0001), Ca2+ amplitude (F = 4.04; p < 0.05), Ca2+ area (F = 0.17; p > 0.05). C, Representative IAHP (arrows) recorded in voltage clamp (top) and corresponding Ca2+ responses (bottom) from four different groups: VP neurons under control conditions (left), VP neurons after 1 h TG incubation, OT neurons under control conditions, and OT neurons after 1 h of TG incubation. Note the inhibited AHP in the TG-incubated VP neuron (asterisk) but lack of effect in OT neurons. D, Summary data of TG's effects on voltage clamp IAHP tail currents evoked by a 30-spike 20 s pulse train from a −55 mV holding potential. Control (VP) n = 13, TG (VP) n = 11, control (OT) n = 14, TG (OT) n = 6. Statistics performed with a one-way ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. 0.1 s amplitude (F = 30.6; p < 0.0001), 1 s amplitude (F = 15.7; p < 0.001), duration (F = 20.9; p < 0.0001), area (F = 24.0; p < 0.0001), Ca2+ amplitude (F = 1.08; p > 0.05), Ca2+ area (F = 2.96; p > 0.05). E, Representative spike trains evoked by a 70 pA depolarizing step under control conditions or after TG incubation. Note the increased spike frequency in the TG-treated neuron. F, Summary data of f-I relationships comparing VP neurons under control versus TG conditions. Control (VP) n = 5, TG (VP) n = 12. Statistics performed with a two-way ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. Spike count [F = 46.6; Control (VP) vs TG (VP) p < 0.0001), spike count (F = 51.8; Control (VP) vs TG (VP) p < 0.0001]. Neuronal gain (f-I slope) is calculated via linear regression significance [F = 5.5; Control (VP) 0.36 ± 0.04 vs TG (VP) 0. 19 ± 0.04; p < 0.05].
Thapsigargin incubation inhibits VP neuron AHPs similarly in male and female rats. AHP parameters of 1 h TG incubation in male and female rats. Male control n = 8; male TG n = 17; female control n = 12; female TG n = 23. Statistics performed with a two-way ANOVA and Šídák's multiple comparisons test. P-values for the overall group comparison are provided in parentheses, for multiple comparisons refer to the figure. 0.1 s amplitude (F = 10.26, p < 0.0001), 1 s amplitude (F = 13.00, p < 0.0001), duration (F = 10.73, p < 0.0001), area (F = 12.20, p < 0.01), Ca2+ peak (F = 6.49, p < 0.01), Ca2+ area (F = 4.38, p = 0.01). Download Figure 1-1, TIF file (5.3MB, tif) .
To further confirm an internal source of Ca2+ from the ER, we generated whole AHPs using photolytic uncaging of Ca2+ using DM-Nitrophen in VP neurons. This bypasses the need for Ca2+ entry via VGCCs during depolarization. As shown in Figure 2, Ca2+ uncaging efficiently evoked a large increase in [Ca2+]c as well as a sAHP. TG preincubation still significantly blocked most of the sAHP component, despite robust Ca2+ increases in response to Ca2+ uncaging (Fig. 2A,B). It should be acknowledged that Ca2+ uncaging may activate other Ca2+ dependent currents such as Ca2+-dependent Cl− channels (Paik et al., 2020), but we argue that contributions from other Ca2+-dependent phenomena are likely minimal due to similar time course of the uncaged response, consistent with sAHPs generated by current injection. Further, the response is inhibited by thapsigargin in a similar manner. The Ca2+ uncaging internal solution does differ significantly from the standard internal solution used in other experiments to accommodate for the properties of DM-Nitrophen. This notably includes the use of an alternative Ca2+ indicator with a different Kd (fluo-4). For these reasons, we retested action potential-evoked (30 spike 20 Hz spike trains) sAHPs and Δ[Ca2+]c using the uncaging internal solution. Similar to our internal solution using fluo-5F, we found that cells preincubated with TG demonstrate significantly smaller sAHP components while Ca2+ transients remain unchanged compared with control conditions (Fig. 2C,D). These results demonstrate that large [Ca2+]c increases can evoke sAHPs without activating VGCCs and that these sAHPs were significantly inhibited by depletion of ER stores.
Figure 2.
TG block of VP AHPs persists when AHPs are evoked by photolytic Ca2+ uncaging. A, Representative example of a Ca2+ trace (top) evoked by a single 50 ms pulse of 405 nm light (blue arrow) and the corresponding current clamp AHP evoked in response to the Ca2+ uncaging (bottom) under control conditions (left) and after TG incubation (right). B, Summary data of TG's effects on AHPs evoked by Ca2+ uncaging. Control n = 11, TG n = 8. 0.1 s amplitude (Mann–Whitney test; p < 0.5), 1 s amplitude (Mann–Whitney test; p < 0.05), duration (unpaired t test; t = 2.47; p < 0.05), area (unpaired t test; t = 2.27; p < 0.05), Ca2+ amplitude (unpaired t test; t = 0.01; p > 0.05), Ca2+ area (Mann–Whitney test; p > 0.05). C, Replication of current clamp AHPs evoked by 20 spike 20 Hz depolarizing pulse trains and corresponding Ca2+ responses using the modified Ca2+ uncaging internal solution containing Ca2+, a different Ca2+ indicator (fluo-4), and DM-Nitrophen which acts as a Ca2+ buffer. TG's effect on AHPs persists under these conditions. D, Summary data of TG's effects on AHPs evoked by 20-spike 20 Hz depolarizing pulse trains using the Ca2+ uncaging internal. Control n = 11, TG n = 8. 0.1 s amplitude (unpaired t test; t = 2.02; p > 0.5), 1 s amplitude (unpaired t test; t = 3.56; p < 0.01), duration (Mann–Whitney test; p < 0.0001), area (unpaired t test; t = 2.26; p < 0.05), Ca2+ amplitude (unpaired t test; t = 1.21; p > 0.05), Ca2+ area (unpaired t test; t = 1.60; p > 0.05).
ER Ca2+ release evoked by caffeine potentiates sAHPs in VP neurons
In addition to ER Ca2+ stores depletion, we assessed the role of ER Ca2+ release in sAHP activation by testing the effect of the ryanodine receptor (RyR) agonist caffeine (10 mM). Caffeine bath application failed to significantly increase the amplitude or the area of the whole AHP waveform generated by pulse trains (Fig. 3A,B). However, we observed a slight but statistically significant increase in the duration of the AHP. The Ca2+ peak was significantly smaller despite an increase in Ca2+ area (Fig. 3B). Additionally, measured input resistance showed a nonsignificant decrease in response to caffeine (control vs caffeine, 880.2 ± 197.1 vs 618.7 ± 126.6 MΩ; p = 0.063, Wilcoxon test). Interestingly, despite the small changes evoked by caffeine on the sAHP, we observed a significant decrease in spike frequency under this condition without a corresponding decrease in neuronal gain (f-I slope; Fig. 3C,D). Moreover, normalizing the amplitude of the AHP by the number of evoked spikes (AHP 0.1 s amplitude/spike count) revealed a significant increase in this AHP/spike index at lower depolarizing steps (10–50 pA) after caffeine application (Fig. 3D, right panel). Despite a general decrease in input resistance, the larger AHP per spike under the caffeine condition supports a genuine effect of caffeine on the AHP. Interestingly, Ca2+ amplitude decreases but cannot be attributed to saturation of the Ca2+ dye, as there is no indication that the signal plateaus. These results indicate that caffeine potentiates AHP activation via ER Ca2+ release following ryanodine receptor activation, leading to decreased firing rates (Fig. 3C).
Figure 3.
Caffeine bath application potentiates AHPs. A, Representative example of a current-clamp 30-spike 20 Hz AHP (top) and the corresponding Ca2+ responses (bottom) before and after caffeine. Current-clamp AHP evoked in response to the uncaging (bottom) under control conditions (left) and after TG incubation (right). B, Summary data of caffeine's effects on current clamp AHPs. Control and caffeine n = 8. 0.1 s amplitude (paired t test; t = 1.87; p > 0.5), 1 s amplitude (paired t test; t = 1.03; p > 0.05), duration (paired t test; t = 2.86; p < 0.05), area (Wilcoxon test; p > 0.05), Ca2+ peak (paired t test; t = 3.64; p < 0.01), Ca2+ area (paired t test; t = 2.52; p < 0.05). C, Representative spike trains evoked by a 70 pA depolarizing step before and after caffeine bath application. Note the decreased spike frequency in the neuron after caffeine. D, Summary data of AHPs before and after caffeine evoked by depolarizing steps. Control and caffeine n = 8. Statistics performed with a two-way ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. Spike Hz (F = 5.2; p < 0.5), AHP amplitude (F = 0.87; p > 0.5), AHP amplitude/spike count (F = 2.21; p = 0.05). Neuronal gain did not significantly differ between groups (control 0.16 ± 0.018 vs caffeine 0.15 ± 0.017; linear regression, F = 0.3, p > 0.05).
Blocking mitochondrial Ca2+ uptake enhances sAHPs
We next tested whether mitochondria, another critical organelle in intracellular Ca2+ handling, shapes the amplitude and time course of the sAHP here, as it has been observed in other cell types (Groten and MacVicar, 2022). Bath application of the mitochondrial uncoupler CCCP (1 µM), which disables mitochondrial via oxidative phosphorylation uncoupling and subsequent collapsing of the mitochondrial membrane potential (Miyazono et al., 2018), significantly enhanced all IsAHP parameters along with corresponding increases in Ca2+ area (Fig. 4A,B). Interestingly, Ca2+ amplitude was slightly but consistently inhibited, suggesting a similar phenomenon to the caffeine effect wherein increased basal Ca2+ lowers the overall amplitude of the Ca2+ waveform. The overall enhancement of the Ca2+ area demonstrates that disabling mitochondria effectively restricted Ca2+ buffering. To further confirm that enhancements in IsAHP reflected a lack of mitochondrial cytosolic Ca2+ buffering, we repeated this experiment with the more selective mitochondrial Ca2+ uniporter (MCU) blocker, Ru360 (20 µM) delivered through the patch pipette (Nathan and Wilson, 2017; Woods et al., 2019; Groten and MacVicar, 2022). After achieving whole-cell configuration, IsAHP values were measured at 5, 10, and 15 min after break-in to track IsAHP parameters in response to progressive Ru360 dialysis. Ru360 significantly enhanced IsAHP amplitude, duration, and area over a 15 min recording period (Fig. 4C,D). Ca2+ amplitude consistently shrank over the 15 min, but this did not reach statistical significance, indicating once again that basal cytosolic Ca2+ is likely increasing. Ca2+ area significantly enhanced over this recording period, suggesting that Ca2+ buffering by mitochondria was impaired. Furthermore, subsequent bath application of CCCP (2 min) failed to further alter any of the IsAHP or Ca2+ measurements (Fig. 4D), indicating that Ru360 dialysis occluded the CCCP effect, confirming in turn that the observed effect of CCCP alone (Fig. 4A) was due to Ca2+ buffering inhibition. Overall, these results indicate that mitochondria Ca2+ buffering normally restricts the IsAHP time course in VP neurons, playing a significant role in shaping the AHP in these neurons.
Figure 4.
Blocking mitochondrial Ca2+ buffering enhances IAHP. A, Representative voltage-clamp trace of IAHP (left) and corresponding Ca2+ responses (right) before and after CCCP bath application. B, Summary data of CCCP's effects on IAHP. Control and CCCP n = 7. 0.1 s amplitude (paired t test; t = 2.54; p < 0.5), 1 s amplitude (paired t test; t = 5.42; p < 0.01), duration (paired t test; t = 5.79; p < 0.01), area (Wilcoxon test; p < 0.05), Ca2+ peak (paired t test; t = 2.61; p < 0.05), Ca2+ area (paired t test; t = 4.40; p < 0.01). C, Representative voltage-clamp trace of IAHP (left) and corresponding Ca2+ responses (right) during Ru360 perfusion through the patch pipette at 5, 10, and 15 min and then subsequent bath application of CCCP. D, Summary data of Ru360 + CCCP's effects on IAHP. n = 10. Statistics performed with a repeated-measures one-way ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. 0.1 s amplitude (F = 11.4; p < 0.01), 1 s amplitude (F = 6.59; p < 0.05), duration (F = 20.51; p < 0.0001), area (F = 8.75; p < 0.01), Ca2+ peak (F = 5.27; p < 0.05), Ca2+ area (F = 9.35; p < 0.01).
Next, we determined whether blocking mitochondrial Ca2+ buffering affected sAHP time course in current-clamp mode and if so, if this blockade translated into slowing of repetitive firing properties. Dialysis of Ru360 through the patch pipette caused a significant enhancement of sAHPs generated by pulse trains over 9 min (Fig. 5A,B). This corresponded with a significant Ca2+ amplitude decrease and Ca2+ area enhancement, consistent with observations in voltage clamp (Fig. 4). Additionally, spiking during depolarizing steps significantly slowed after 9 min of Ru360 dialysis, compared with initial spike frequency measurements (Fig. 5C,D). These data suggest that not only do mitochondria play an important role in shaping AHP time course but that they efficiently influence repetitive firing frequency properties VP neurons.
Figure 5.
Blocking the mitochondrial Ca2+ uniporter with Ru360 enhances the AHP and slows firing in current clamp. A, Representative AHPs evoked with 30 spikes at 20 Hz (left) and corresponding Ca2+ responses (right) at 1 and 9 min after break-in. Note the enhancement of the AHP and Ca2+ signal time course after 9 min of perfusion. B, Summary data of Ru360 effects on AHPs compared with neurons without Ru360. Control n = 4, Ru360 n = 6. Statistics performed with a repeated-measures two-way ANOVA and Šídák's multiple-comparisons test. p values for the group comparison are provided in parentheses; for multiple comparisons, refer to the figure. 0.1 s amplitude (F = 28.72; p < 0.0001), 1 s amplitude (F = 14.52; p < 0.0001), duration (F = 18.86; p < 0.0001), area (F = 10.37; p < 0.0001), Ca2+ peak (F = 16.48; p < 0.0001), Ca2+ area (F = 13.69; p < 0.0001). C, Representative example of spiking in response to a 70 pA depolarizing step 1 min and 9 min after achieving whole-cell configuration with Ru360 in the pipette. Note the lower spike frequency at 9 min. Summary data of Ru360 effects on AHP spike frequency performed with a one-way ANOVA and Šídák's multiple-comparisons test (n = 5; F = 21.76; p = 0.001). For multiple comparisons refer to the figure.
Block of mitochondrial Ca2+ efflux does not significantly affect sAHPs
In addition to Ca2+ buffering via the MCU, mitochondria also contain pumps that can exude Ca2+ into the cytosol in a Na+-dependent or Na+-independent manner (Bernardi, 1999). Since mitochondrial buffering specifically modulates sAHP time course, we wanted to test whether blocking mitochondrial Ca2+ efflux would inhibit sAHPs. TPP+ is a compound that specifically inhibits mitochondrial Ca2+ efflux (Karadjov et al., 1986). It should be noted that TPP+ fully blocks Na+-dependent Ca2+ efflux while blocking ∼50% of Na+-independent Ca2+ efflux at the concentration used (10 µM; Wingrove and Gunter, 1986). Attempts at higher concentrations of TPP+ produced inconsistent results and often compromised cell viability in our preparation, so we used 10 µM despite incomplete block. Bath application of TPP+ failed to inhibit the AHP in either current clamp (Fig. 6A,B) or voltage clamp (Fig. 6C,D). Corresponding Ca2+ peak and areas after TPP+ application remained unchanged except that Ca2+ area increased in current clamp (Fig. 6B). Subsequent application of CCCP (1 µM) produced significant enhancements to AHP and IAHP duration and area (Fig. 6B,D) as previously observed. Corresponding Ca2+ area, but not peak, significantly increased in both current and voltage clamp (Fig. 6B,D). Together, these results indicate that in stark contrast to mitochondrial Ca2+ uptake, mitochondrial Ca2+ efflux plays no consequential role in sAHP activation or modulation. To investigate the possibility that these compounds may be increasing the overall baseline Ca2+, we compared raw baseline Ca2+ signals from this data in both voltage and current clamp and found no significant differences between the three groups (Extended Data Fig. 6-1).
Figure 6.
Blocking mitochondrial Ca2+ pumps via TPP+ does not significantly effect AHPs. A, Representative example of TPP+'s effect on AHPs (top) and corresponding Ca2+ responses (bottom). Note that while TPP+ has little effect, subsequent application of CCCP enhances the AHP. See Extended Data Figure 6-1 for an analysis of baseline Ca2+ differences between the groups. B, Summary data of TPP+ effects on AHPs. n = 8. Statistics performed with a one-way repeated-measures ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses; for multiple comparisons, refer to the figure. 0.1 s amplitude (F = 3.14; p > 0.05), 1 s amplitude (F = 6.18; p < 0. 05), duration (F = 22.51; p < 0.01), area (F = 8.01; p < 0.05), Ca2+ peak (F = 0.34; p > 0.05), Ca2+ area (F = 9.78; p < 0.05). C, Representative example of TPP+'s effect on IAHP (top) and corresponding Ca2+ responses (bottom). Note that while TPP+ has little effect, subsequent application of CCCP enhances the AHP. D, Summary data of TPP+ effects on IAHP. n = 11. Statistics performed with a one-way repeated-measures ANOVA and Šídák's multiple-comparisons test. p values for the overall test are provided in parentheses, for multiple comparisons refer to the figure. 0.1 s amplitude (F = 6.74; p < 0.05), 1 s amplitude (F = 6.32; p < 0.05), duration (F = 7.98; p < 0.05), area (F = 9.25; p < 0.01), Ca2+ peak (F = 3.79; p > 0.05), Ca2+ area (F = 10.38; p < 0.01).
Baseline Ca2+ measurements do not significantly differ after TPP+ or subsequent CCCP. Raw baseline fluorescence for Control, TPP+, and CCCP in both voltage clamp (left) and current clamp (right). In voltage clamp, we observed a significant difference (One-way repeated measures ANOVA, F = 4.80, p = 0.031) however post hoc analysis revealed no significant between group differences. In current clamp, we observed a significant difference (One-way repeated measures ANOVA, F = 4.11, p = 0.042) however post hoc analysis revealed no significant between group differences. Download Figure 6-1, TIF file (2.3MB, tif) .
Discussion
This study was designed to determine a Ca2+ source activating sAHPs in VP neurons. Previous work demonstrated N-type channels were primarily responsible for activating the mAHP and sAHP in OT neurons while coupling of specific VGCCs to the mAHP and sAHP in VP neurons was less clear (Kirchner et al., 2018). The largest contributions were a 29% reduction after N-type block via conotoxin GVIA and a 25% reduction in sAHPs after R-type block via SNX-482. Thus, no single Ca2+ source accounted for a majority of the sAHP in VP neurons, compared with a more comprehensive block of the OT AHP components by conotoxin GVIA (Kirchner et al., 2018). We therefore tested the contribution of ER Ca2+ stores to sAHP activation in VP neurons.
We show that depletion of ER stores via TG abolishes the sAHP, suggesting that ER Ca2+ plays a major role activating the sAHP in VP neurons (Fig. 1). Furthermore, RyR activation by caffeine appears to potentiate sAHP activation (Fig. 3) by a mechanism yet to be determined. This is similar to ryanodine-dependent sAHP activation described in other neuronal cell types (Pineda et al., 1999; Shah and Haylett, 2000; van de Vrede et al., 2007; Sahu et al., 2019) and overall suggests the sAHP in VP neurons relies heavily on Ca2+-induced ER Ca2+ release (CICR). One parsimonious explanation for the major contributors is that R-type and N-type channels provide Ca2+ that activates CICR from ER. Coupling between N-type VGCCs and ER in MNCs is previously established within the context of somatodendritic peptide release, a unique phenomenon of MNCs wherein Ca2+-dependent exocytosis of OT and VP occurs in the soma and dendrites (Ludwig, 1998; de Kock et al., 2003; Tobin et al., 2004; Brown et al., 2020). Thapsigargin depletion mobilizes large dense core vesicles (LDCVs) to readily releasable pools at plasma membrane (Ludwig et al., 2002; Tobin et al., 2004), and this mobilization is blunted in the presence of N-type blocker ω-conotoxin GVIA. Moreover, the N-type channel has high association to Golgi apparatus where peptide is packaged into LDCVs (Tobin et al., 2011). With regard to R-type VGCCs, close association with ER is less established but they interestingly show much higher expression in VP neurons compared with OT (Tobin et al., 2011). The significance of R-type channels in VP MNC neurophysiology requires further study.
The resultant [Ca2+]c inhibition by VGCC blockade does not correspond with a proportional inhibition of the sAHP, suggesting that it is likely activated by a Ca2+ microdomain instead of reflecting the magnitude and temporal trajectory of a global [Ca2+]c rise (regardless of the source). Evidence for this sort of proportionality exists in neocortical pyramidal neurons, where there is a sigmoidal cooperative relationship between IsAHP amplitude and [Ca2+]c (Abel et al., 2004). This does not appear to be the case in VP neurons, where we found that depletion of ER Ca2+ stores significantly reduces sAHPs (69–92%) without meaningfully reducing global [Ca2+]c (Figs. 1, 2). This suggests that measurements of global [Ca2+]c are a poor indicator of the properties of the evoked sAHP, arguing also against its contribution as an underlying mechanism triggering its activation. Global [Ca2+]c increases are still critical to the activation of the sAHP, evidenced by its complete blockade when Δ[Ca2+]c are prevented (Ghamari-Langroudi and Bourque, 2004; Kirchner et al., 2018). Rather, we propose that a Ca2+ threshold must be met at specific sites to activate relevant signaling cascades. It appears sAHP signaling activated by ER Ca2+ is mediated by a microdomain that is modulated without an obvious impact in global [Ca2+]c. This is consistent with previous data in VP neurons showing that [Ca2+]c decay is dissociated from AHP decay, suggesting compartmentalization of AHP-activating Ca2+ (Roper et al., 2003).
The significance underlying the contrasting sAHP Ca2+ sources of SON OT and VP neurons is at present unknown. Despite their indiscriminate morphological appearance, these neurons have considerably different electrophysiological properties (Armstrong and Stern, 1998; Armstrong et al., 2019), including differences in their dependence on PIP2 to activate sAHPs and VGGCs (Kirchner et al., 2017, 2019). Thus, it is tempting to speculate that differences in sAHP Ca2+ sources could be associated to the differing physiological roles OT and VP play within the organism and their secretion in response to differing stimuli. An instructive example of this is the AHP plasticity that occurs in OT but not VP neurons during pregnancy and lactation, where it is hypothesized that this AHP enhancement helps facilitate the short, high-frequency OT milk ejection bursts (Stern and Armstrong, 1996; Teruyama and Armstrong, 2002, 2005; Wang et al., 2018). OT neurons can coordinate their activity and burst synchronously while VP neurons tend to demonstrate asynchronous phasic firing, which results in sustained systemic secretion of vasopressin needed to cope with an osmotic challenge (Bourque et al., 1998). Another notable difference between OT and VP neurons is their calbindin and calretinin expression levels. Previous studies report both these Ca2+-buffering proteins express in 70% of OT neurons while both are absent in 73% of VP neurons (Arai et al., 1999). This suggests that VP neurons have a lower overall Ca2+ buffering capacity compared with OT neurons, which could partially explain the role of CICR. Furthermore, this is consistent with reports that VP neurons generally have larger AHPs, though this difference is subtle (Armstrong et al., 1994; Stern and Armstrong, 1996; Teruyama and Armstrong, 2002). Clearly, future studies are needed to further understand the significance of the different Ca2+ sources triggering AHPs in VP and OT neurons.
We also tested the contribution of the mitochondria to sAHPs in VP neurons by either disabling them completely (Fig. 4), blocking the Ca2+ uptake uniporter MCU (Figs. 4, 5), or disabling the Na+–Ca2+ exchanger (Fig. 6). Blocking mitochondria Ca2+ uptake significantly enhanced the sAHP, while blocking Ca2+ extrusion had no effect, indicating that mitochondria does not contribute as a Ca2+ source to activate that sAHP but that it rather restricts the sAHP time course by actively buffering Ca2+. This is in line with a previous study showing that the primary Ca2+ handling role of mitochondria is to buffer it in VP neurons (Komori et al., 2010).
ER and mitochondria membranes can form tight complexes called mitochondria-associated membranes (MAMs; Rizzuto et al., 1998). These MAMs are dynamic protein scaffolds wherein high spatiotemporal control of local Ca2+ microdomains can coordinate signaling pathways (Patergnani et al., 2011; Giorgi et al., 2015; Missiroli et al., 2018). MAMs contain Ca2+ sensors, which are critically important for sAHP activation in other cell types, including hippocalcin in CA1 (Tzingounis et al., 2007), neurocalcin-δ cortical pyramidal neurons (Villalobos and Andrade, 2010), and calcineurin in myenteric neurons (Vogalis et al., 2004). MAMs may play an important role in providing a spatially constricted, microdomain avenue for sAHP activation in VP neurons. This is supported by our data showing a high efficiency of mitochondrial Ca2+ clearing combined with ER Ca2+ release to dictate the activation and spatiotemporal dynamics of the sAHP. These interactions are critical for not only understanding intrinsic excitability, but also VP somatodendritic release, as ER Ca2+ critically primes VP-containing LDCVs to the plasma membrane (Li and Hatton, 1997a; Tobin et al., 2004; Tobin and Ludwig, 2007; Kortus et al., 2016). In fact, given the strong overlap in Ca2+ sources by both the sAHP and somatodendritic release, it is tempting to speculate that sAHP may play a role in modulating the volume and pattern of somatodendritic release. Experiments to test this possibility as well as determining whether mitochondrial Ca2+ buffering plays a role in somatodendritic release are warranted.
The sAHP is one of many intrinsic excitability mechanisms that shape the phasic firing patterns of VP neurons (Thirouin and Bourque, 2021). During MNC burst onset, the reciprocal depolarizing afterpotential (DAP) activates, sustaining a plateau potential for phasic burst onset and regeneration via CICR activation (Andrew and Dudek, 1984b; Armstrong et al., 1994; Li and Hatton, 1997a,b). This present study suggests that the ER in VP neurons plays a role in regulating the intrinsic excitability of VP neurons by activating both DAP and sAHP, likely through a differing spatiotemporal pattern of ER Ca2+ release during the burst phase.
The sAHP is a mechanism that is commonly affected in many disease states where neuronal firing is abnormal. For example, cortical pyramidal neurons in epileptic subjects have diminished sAHPs, (Lorenzon and Foehring, 1992; Soh and Tzingounis, 2010; Tiwari et al., 2022) and neurons from Alzheimer's mouse models show enhanced sAHPs (Landfield and Pitler, 1984; Moyer et al., 1992; Power et al., 2002; Kaczorowski et al., 2011). Furthermore, previous work from our laboratory demonstrated hyperexcitability of VP neurons in a rat model of heart failure (Han et al., 2010; Potapenko et al., 2011, 2013; Stern and Potapenko, 2013; Ferreira-Neto et al., 2017; Ferreira-Neto and Stern, 2019), which contributes to neurohumoral activation, morbidity, and mortality in this prevalent disease (Goldsmith et al., 1983; Francis et al., 1990; Chatterjee, 2005). In this context, we demonstrated that a blunted mAHP current despite normal somatic Ca2+ elevation in response to pulse trains constitutes an underlying mechanism leading to VP hyperexcitability in this disease (Ferreira-Neto et al., 2017). Thus, future studies to determine whether the sAHP in VP neurons and the underlying ER CICR mechanism are also affected during heart failure are warranted.
References
- Abel HJ, Lee JCF, Callaway JC, Foehring RC (2004) Relationships between intracellular calcium and afterhyperpolarizations in neocortical pyramidal neurons. J Neurophysiol 91:324–335. 10.1152/jn.00583.2003 [DOI] [PubMed] [Google Scholar]
- Adelman JP, Maylie J, Sah P (2012) Small-conductance Ca2+-activated K+ channels: form and function. Annu Rev Physiol 74:245–269. 10.1146/annurev-physiol-020911-153336 [DOI] [PubMed] [Google Scholar]
- Andrade R, Foehring RC, Tzingounis AV (2012) The calcium-activated slow AHP: cutting through the Gordian knot. Front Cell Neurosci 6:47. 10.3389/fncel.2012.00047 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrew RD, Dudek FE (1984a) Intrinsic inhibition in magnocellular neuroendocrine cells of rat hypothalamus. J Physiol 353:171–185. 10.1113/jphysiol.1984.sp015330 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrew RD, Dudek FE (1984b) Analysis of intracellularly recorded phasic bursting by mammalian neuroendocrine cells. J Neurophysiol 51:552–566. 10.1152/jn.1984.51.3.552 [DOI] [PubMed] [Google Scholar]
- Arai R, Jacobowitz DM, Hida T (1999) Calbindin D28k and calretinin in oxytocin and vasopressin neurons of the rat supraoptic nucleus. Cell Tissue Res 298:11–19. 10.1007/PL00008808 [DOI] [PubMed] [Google Scholar]
- Armstrong WE, Foehring RC, Kirchner MK, Sladek CD (2019) Electrophysiological properties of identified oxytocin and vasopressin neurones. J Neuroendocrinol 31:e12666. 10.1111/jne.12666 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armstrong WE, Gallagher MJ, Sladek CD (1986) Noradrenergic stimulation of supraoptic neuronal activity and vasopressin release in vitro: mediation by an α1-receptor. Brain Res 365:192–197. 10.1016/0006-8993(86)90739-0 [DOI] [PubMed] [Google Scholar]
- Armstrong WE, Smith BN, Tian M (1994) Electrophysiological characteristics of immunochemically identified rat oxytocin and vasopressin neurones in vitro. J Physiol 475:115–128. 10.1113/jphysiol.1994.sp020053 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armstrong WE, Stern JE (1998) Electrophysiological distinctions between oxytocin and vasopressin neurons in the supraoptic nucleus. Adv Exp Med Biol 449:67–77. 10.1007/978-1-4615-4871-3_7 [DOI] [PubMed] [Google Scholar]
- Armstrong WE, Wang L, Li C, Teruyama R (2010) Performance, properties and plasticity of identified oxytocin and vasopressin neurones in vitro. J Neuroendocrinol 22:330–342. 10.1111/j.1365-2826.2010.01989.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bernardi P (1999) Mitochondrial transport of cations: channels, exchangers, and permeability transition. Physiol Rev 79:1127–1155. 10.1152/physrev.1999.79.4.1127 [DOI] [PubMed] [Google Scholar]
- Bicknell RJ (1988) Optimizing release from peptide hormone secretory nerve terminals. J Exp Biol 139:51–65. 10.1242/jeb.139.1.51 [DOI] [PubMed] [Google Scholar]
- Bicknell RJ, Leng G (1981) Relative efficiency of neural firing patterns for vasopressin release in vitro. Neuroendocrinology 33:295–299. 10.1159/000123248 [DOI] [PubMed] [Google Scholar]
- Bourque CW, Brown DA (1987) Apamin and d-tubocurarine block the after-hyperpolarization of rat supraoptic neurosecretory neurons. Neurosci Lett 82:185–190. 10.1016/0304-3940(87)90127-3 [DOI] [PubMed] [Google Scholar]
- Bourque CW, Kirkpatrick K, Jarvis CR (1998) Extrinsic modulation of spike afterpotentials in rat hypothalamoneurohypophysial neurons. Cell Mol Neurobiol 18:3–12. 10.1023/A:1022566924921 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bourque CW, Randle JC, Renaud LP (1985) Calcium-dependent potassium conductance in rat supraoptic nucleus neurosecretory neurons. J Neurophysiol 54:1375–1382. 10.1152/jn.1985.54.6.1375 [DOI] [PubMed] [Google Scholar]
- Brown CH, Ludwig M, Tasker JG, Stern JE (2020) Somato-dendritic vasopressin and oxytocin secretion in endocrine and autonomic regulation. J Neuroendocrinol 32:e12856. 10.1111/jne.12856 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cazalis M, Dayanithi G, Nordmann JJ (1985) The role of patterned burst and interburst interval on the excitation-coupling mechanism in the isolated rat neural lobe. J Physiol 369:45–60. 10.1113/jphysiol.1985.sp015887 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chatterjee K (2005) Neurohormonal activation in congestive heart failure and the role of vasopressin. Am J Cardiol 95:8B–13B. 10.1016/j.amjcard.2005.03.003 [DOI] [PubMed] [Google Scholar]
- Dayanithi G, Forostyak O, Ueta Y, Verkhratsky A, Toescu EC (2012) Segregation of calcium signalling mechanisms in magnocellular neurones and terminals. Cell Calcium 51:293–299. 10.1016/j.ceca.2012.02.002 [DOI] [PubMed] [Google Scholar]
- de Kock CPJ, Wierda KDB, Bosman LWJ, Min R, Koksma J-J, Mansvelder HD, Verhage M, Brussaard AB (2003) Somatodendritic secretion in oxytocin neurons is upregulated during the female reproductive cycle. J Neurosci 23:2726–2734. 10.1523/JNEUROSCI.23-07-02726.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dopico AM, Widmer H, Wang G, Lemos JR, Treistman SN (1999) Rat supraoptic magnocellular neurones show distinct large conductance, Ca2+-activated K+ channel subtypes in cell bodies versus nerve endings. J Physiol 519(Pt 1):101–114. 10.1111/j.1469-7793.1999.0101o.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dutton A, Dyball REJ (1979) Phasic firing enhances vasopressin release from the rat neurohypophysis. J Physiol 290:433–440. 10.1113/jphysiol.1979.sp012781 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erickson KR, Ronnekleiv OK, Kelly MJ (1993) Role of a T-type calcium current in supporting a depolarizing potential, damped oscillations, and phasic firing in vasopressinergic Guinea pig supraoptic neurons. Neuroendocrinology 57:789–800. 10.1159/000126438 [DOI] [PubMed] [Google Scholar]
- Faas GC, Karacs K, Vergara JL, Mody I (2005) Kinetic properties of DM-nitrophen binding to calcium and magnesium. Biophys J 88:4421–4433. 10.1529/biophysj.104.057745 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferreira-Neto HC, Biancardi VC, Stern JE (2017) A reduction in SK channels contributes to increased activity of hypothalamic magnocellular neurons during heart failure. J Physiol 595:6429–6442. 10.1113/JP274730 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferreira-Neto HC, Stern JE (2019) Functional coupling between NMDA receptors and SK channels in rat hypothalamic magnocellular neurons: altered mechanisms during heart failure. J Physiol 599:507–520. 10.1113/JP278910 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Francis GS, Benedict C, Johnstone DE, Kirlin PC, Nicklas J, Liang CS, Kubo SH, Rudin-Toretsky E, Yusuf S (1990) Comparison of neuroendocrine activation in patients with left ventricular dysfunction with and without congestive heart failure. A substudy of the studies of left ventricular dysfunction (SOLVD). Circulation 82:1724–1729. 10.1161/01.CIR.82.5.1724 [DOI] [PubMed] [Google Scholar]
- Gainer H, Wolfe SA, Obaid AL, Salzberg BM (1986) Action potentials and frequency-dependent secretion in the mouse neurohypophysis. Neuroendocrinology 43:557–563. 10.1159/000124582 [DOI] [PubMed] [Google Scholar]
- Ghamari-Langroudi M, Bourque CW (1998) Caesium blocks depolarizing after-potentials and phasic firing in rat supraoptic neurones. J Physiol 510:165–175. 10.1111/j.1469-7793.1998.165bz.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghamari-Langroudi M, Bourque CW (2004) Muscarinic receptor modulation of slow afterhyperpolarization and phasic firing in rat supraoptic nucleus neurons. J Neurosci 24:7718–7726. 10.1523/JNEUROSCI.1240-04.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giorgi C, Missiroli S, Patergnani S, Duszynski J, Wieckowski MR, Pinton P (2015) Mitochondria-associated membranes: composition, molecular mechanisms, and physiopathological implications. Antioxid Redox Signal 22:995–1019. 10.1089/ars.2014.6223 [DOI] [PubMed] [Google Scholar]
- Goldsmith SR, Francis GS, Cowley AW, Levine TB, Cohn JN (1983) Increased plasma arginine vasopressin levels in patients with congestive heart failure. J Am Coll Cardiol 1:1385–1390. 10.1016/S0735-1097(83)80040-0 [DOI] [PubMed] [Google Scholar]
- Greffrath W, Martin E, Reuss S, Boehmer G (1998) Components of after-hyperpolarization in magnocellular neurones of the rat supraoptic nucleus in vitro. J Physiol 513:493–506. 10.1111/j.1469-7793.1998.493bb.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Groten CJ, MacVicar BA (2022) Mitochondrial Ca2+ uptake by the MCU facilitates pyramidal neuron excitability and metabolism during action potential firing. Commun Biol 5:900. 10.1038/s42003-022-03848-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hajnóczky G, Csordás G, Yi M (2002) Old players in a new role: mitochondria-associated membranes, VDAC, and ryanodine receptors as contributors to calcium signal propagation from endoplasmic reticulum to the mitochondria. Cell Calcium 32:363–377. 10.1016/S0143416002001872 [DOI] [PubMed] [Google Scholar]
- Han TH, Lee K, Park JB, Ahn D, Park J-H, Kim D-Y, Stern JE, Lee SY, Ryu PD (2010) Reduction in synaptic GABA release contributes to target-selective elevation of PVN neuronal activity in rats with myocardial infarction. Am J Physiol Regul Integr Comp Physiol 299:R129–R139. 10.1152/ajpregu.00391.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaczorowski CC, Sametsky E, Shah S, Vassar R, Disterhoft JF (2011) Mechanisms underlying basal and learning-related intrinsic excitability in a mouse model of Alzheimer’s disease. Neurobiol Aging 32:1452–1465. 10.1016/j.neurobiolaging.2009.09.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karadjov JS, Kudzina LY, Zinchenko VP (1986) TPP+ inhibits Na+-stimulated Ca2+ efflux from brain mitochondria. Cell Calcium 7:115–119. 10.1016/0143-4160(86)90014-X [DOI] [PubMed] [Google Scholar]
- Karagas NE, Venkatachalam K (2019) Roles for the endoplasmic reticulum in regulation of neuronal calcium homeostasis. Cells 8:1232. 10.3390/cells8101232 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim KS, Duignan KM, Hawryluk JM, Soh H, Tzingounis AV (2016) The voltage activation of cortical KCNQ channels depends on global PIP2 levels. Biophys J 110:1089–1098. 10.1016/j.bpj.2016.01.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirchner MK, Armstrong WE, Guan D, Ueta Y, Foehring RC (2019) PIP2 alters Ca2+ currents in acutely dissociated supraoptic oxytocin neurons. Physiol Rep 7:e14198. 10.14814/phy2.14198 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirchner MK, Foehring RC, Callaway JC, Armstrong WE (2018) Specificity in the interaction of high-voltage-activated Ca2+ channel types with Ca2+-dependent afterhyperpolarizations in magnocellular supraoptic neurons. J Neurophysiol 120:1728–1739. 10.1152/jn.00285.2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirchner MK, Foehring RC, Wang L, Chandaka GK, Callaway JC, Armstrong WE (2017) Phosphatidylinositol 4, 5-bisphosphate (PIP2) modulates afterhyperpolarizations in oxytocin neurons of the supraoptic nucleus. J Physiol 595:4927–4946. 10.1113/JP274219 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Komori Y, Tanaka M, Kuba M, Ishii M, Abe M, Kitamura N, Verkhratsky A, Shibuya I, Dayanithi G (2010) Ca2+ homeostasis, Ca2+ signalling and somatodendritic vasopressin release in adult rat supraoptic nucleus neurones. Cell Calcium 48:324–332. 10.1016/j.ceca.2010.10.002 [DOI] [PubMed] [Google Scholar]
- Kortus S, Srinivasan C, Forostyak O, Zapotocky M, Ueta Y, Sykova E, Chvatal A, Verkhratsky A, Dayanithi G (2016) Sodium-calcium exchanger and R-type Ca2+ channels mediate spontaneous [Ca2+]i oscillations in magnocellular neurones of the rat supraoptic nucleus. Cell Calcium 59:289–298. 10.1016/j.ceca.2016.03.010 [DOI] [PubMed] [Google Scholar]
- Lancaster B, Adams PR (1986) Calcium-dependent current generating the afterhyperpolarization of hippocampal neurons. J Neurophysiol 55:1268–1282. 10.1152/jn.1986.55.6.1268 [DOI] [PubMed] [Google Scholar]
- Landfield PW, Pitler TA (1984) Prolonged Ca2+-dependent afterhyperpolarizations in hippocampal neurons of aged rats. Science 226:1089–1092. 10.1126/science.6494926 [DOI] [PubMed] [Google Scholar]
- Li Z, Hatton GI (1997a) Ca2+ release from internal stores: role in generating depolarizing after-potentials in rat supraoptic neurones. J Physiol 498:339–350. 10.1113/jphysiol.1997.sp021862 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Z, Hatton GI (1997b) Reduced outward K+ conductances generate depolarizing after-potentials in rat supraoptic nucleus neurones. J Physiol (Lond) 505:95–106. 10.1111/j.1469-7793.1997.095bc.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lorenzon NM, Foehring RC (1992) Relationship between repetitive firing and afterhyperpolarizations in human neocortical neurons. J Neurophysiol 67:350–363. 10.1152/jn.1992.67.2.350 [DOI] [PubMed] [Google Scholar]
- Ludwig M (1998) Dendritic release of vasopressin and oxytocin. J Neuroendocrinol 10:881–895. 10.1046/j.1365-2826.1998.00279.x [DOI] [PubMed] [Google Scholar]
- Ludwig M, Sabatier N, Bull PM, Landgraf R, Dayanithi G, Leng G (2002) Intracellular calcium stores regulate activity-dependent neuropeptide release from dendrites. Nature 418:85–89. 10.1038/nature00822 [DOI] [PubMed] [Google Scholar]
- McCormick DA, Contreras D (2001) On the cellular and network bases of epileptic seizures. Annu Rev Physiol 63:815–846. 10.1146/annurev.physiol.63.1.815 [DOI] [PubMed] [Google Scholar]
- Missiroli S, Patergnani S, Caroccia N, Pedriali G, Perrone M, Previati M, Wieckowski MR, Giorgi C (2018) Mitochondria-associated membranes (MAMs) and inflammation. Cell Death Dis 9:1–14. 10.1038/s41419-017-0027-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miyazono Y, Hirashima S, Ishihara N, Kusukawa J, Nakamura K, Ohta K (2018) Uncoupled mitochondria quickly shorten along their long axis to form indented spheroids, instead of rings, in a fission-independent manner. Sci Rep 8:350. 10.1038/s41598-017-18582-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moyer JR, Thompson LT, Black JP, Disterhoft JF (1992) Nimodipine increases excitability of rabbit CA1 pyramidal neurons in an age- and concentration-dependent manner. J Neurophysiol 68:2100–2109. 10.1152/jn.1992.68.6.2100 [DOI] [PubMed] [Google Scholar]
- Nathan SR, Wilson JJ (2017) Synthesis and evaluation of a ruthenium-based mitochondrial calcium uptake inhibitor. J Vis Exp 128:56527. 10.3791/56527 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paik S-S, Park YS, Kim I-B (2020) Calcium- and voltage-dependent dual gating ANO1 is an intrinsic determinant of repolarization in rod bipolar cells of the mouse retina. Cells 9:543. 10.3390/cells9030543 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Patergnani S, et al. (2011) Calcium signaling around mitochondria associated membranes (MAMs). Cell Commun Signal 9:19. 10.1186/1478-811X-9-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pineda JC, Galarraga E, Foehring RC (1999) Different Ca2+ source for slow AHP in completely adapting and repetitive firing pyramidal neurons. Neuroreport 10:1951–1956. 10.1097/00001756-199906230-00029 [DOI] [PubMed] [Google Scholar]
- Potapenko ES, Biancardi VC, Florschutz RM, Ryu PD, Stern JE (2011) Inhibitory-excitatory synaptic balance is shifted toward increased excitation in magnocellular neurosecretory cells of heart failure rats. J Neurophysiol 106:1545–1557. 10.1152/jn.00218.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Potapenko ES, Biancardi VC, Zhou Y, Stern JE (2013) Astrocytes modulate a postsynaptic NMDA-GABAA-receptor crosstalk in hypothalamic neurosecretory neurons. J Neurosci 33:631–640. 10.1523/JNEUROSCI.3936-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poulain DA, Wakerley JB (1982) Electrophysiology of hypothalamic magnocellular neurones secreting oxytocin and vasopressin. Neuroscience 7:773–808. 10.1016/0306-4522(82)90044-6 [DOI] [PubMed] [Google Scholar]
- Power JM, Wu WW, Sametsky E, Oh MM, Disterhoft JF (2002) Age-related enhancement of the slow outward calcium-activated potassium current in hippocampal CA1 pyramidal neurons in vitro. J Neurosci 22:7234–7243. 10.1523/JNEUROSCI.22-16-07234.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizzuto R, et al. (2009) Ca2+ transfer from the ER to mitochondria: when, how and why. Biochim Biophys Acta 1787:1342–1351. 10.1016/j.bbabio.2009.03.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizzuto R, Pinton P, Carrington W, Fay FS, Fogarty KE, Lifshitz LM, Tuft RA, Pozzan T (1998) Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses. Science 280:1763–1766. 10.1126/science.280.5370.1763 [DOI] [PubMed] [Google Scholar]
- Rizzuto R, Stefani DD, Raffaello A, Mammucari C (2012) Mitochondria as sensors and regulators of calcium signalling. Nat Rev Mol Cell Biol 13:566–578. 10.1038/nrm3412 [DOI] [PubMed] [Google Scholar]
- Roper P, Callaway J, Shevchenko T, Teruyama R, Armstrong W (2003) AHP’s, HAP’s and DAP’s: how potassium currents regulate the excitability of rat supraoptic neurones. J Comput Neurosci 15:367–389. 10.1023/A:1027424128972 [DOI] [PubMed] [Google Scholar]
- Sah P, McLachlan EM (1991) Ca(2+)-activated K+ currents underlying the afterhyperpolarization in Guinea pig vagal neurons: a role for Ca(2+)-activated Ca2+ release. Neuron 7:257–264. 10.1016/0896-6273(91)90264-Z [DOI] [PubMed] [Google Scholar]
- Sahu G, Wazen R-M, Colarusso P, Chen SRW, Zamponi GW, Turner RW (2019) Junctophilin proteins tether a Cav1-RyR2-KCa3.1 tripartite complex to regulate neuronal excitability. Cell Rep 28:2427–2442.e6. 10.1016/j.celrep.2019.07.075 [DOI] [PubMed] [Google Scholar]
- Sato-Numata K, Numata T, Ueta Y, Okada Y (2020) Expression and functions of N-type Cav2.2 and T-type Cav3.1 channels in rat vasopressin neurons under normotonic conditions. J Physiol Sci 70:49. 10.1186/s12576-020-00775-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwindt PC, Spain WJ, Foehring RC, Chubb MC, Crill WE (1988a) Slow conductances in neurons from cat sensorimotor cortex in vitro and their role in slow excitability changes. J Neurophysiol 59:450–467. 10.1152/jn.1988.59.2.450 [DOI] [PubMed] [Google Scholar]
- Schwindt PC, Spain WJ, Foehring RC, Stafstrom CE, Chubb MC, Crill WE (1988b) Multiple potassium conductances and their functions in neurons from cat sensorimotor cortex in vitro. J Neurophysiol 59:424–449. 10.1152/jn.1988.59.2.424 [DOI] [PubMed] [Google Scholar]
- Shah M, Haylett DG (2000) Ca(2+) channels involved in the generation of the slow afterhyperpolarization in cultured rat hippocampal pyramidal neurons. J Neurophysiol 83:2554–2561. 10.1152/jn.2000.83.5.2554 [DOI] [PubMed] [Google Scholar]
- Soh H, Tzingounis AV (2010) The specific slow afterhyperpolarization inhibitor UCL2077 is a subtype-selective blocker of the epilepsy associated KCNQ channels. Mol Pharmacol 78:1088–1095. 10.1124/mol.110.066100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stern JE, Armstrong WE (1996) Changes in the electrical properties of supraoptic nucleus oxytocin and vasopressin neurons during lactation. J Neurosci 16:4861–4871. 10.1523/JNEUROSCI.16-16-04861.1996 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stern JE, Potapenko ES (2013) Enhanced NMDA receptor-mediated intracellular calcium signaling in magnocellular neurosecretory neurons in heart failure rats. Am J Physiol Regul Integr Comp Physiol 305:R414–R422. 10.1152/ajpregu.00160.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szalai G, Csordás G, Hantash BM, Thomas AP, Hajnóczky G (2000) Calcium signal transmission between ryanodine receptors and mitochondria. J Biol Chem 275:15305–15313. 10.1074/jbc.275.20.15305 [DOI] [PubMed] [Google Scholar]
- Takeuchi A, Kim B, Matsuoka S (2015) The destiny of Ca2+ released by mitochondria. J Physiol Sci 65:11–24. 10.1007/s12576-014-0326-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teruyama R, Armstrong WE (2002) Changes in the active membrane properties of rat supraoptic neurones during pregnancy and lactation. J Neuroendocrinol 14:933–944. 10.1046/j.1365-2826.2002.00844.x [DOI] [PubMed] [Google Scholar]
- Teruyama R, Armstrong WE (2005) Enhancement of calcium-dependent afterpotentials in oxytocin neurons of the rat supraoptic nucleus during lactation. J Physiol 566:505–518. 10.1113/jphysiol.2005.085985 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teruyama R, Lipschitz DL, Wang L, Ramoz GR, Crowley WR, Bealer SL, Armstrong WE (2008) Central blockade of oxytocin receptors during mid-late gestation reduces amplitude of slow afterhyperpolarization in supraoptic oxytocin neurons. Am J Physiol Endocrinol Metab 295:E1167–E1171. 10.1152/ajpendo.90620.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thirouin ZS, Bourque CW (2021) Mechanism and function of phasic firing in vasopressin-releasing magnocellular neurosecretory cells. J Neuroendocrinol 33:e13048. 10.1111/jne.13048 [DOI] [PubMed] [Google Scholar]
- Tiwari MN, Mohan S, Biala Y, Shor O, Benninger F, Yaari Y (2022) Corticotropin releasing factor mediates KCa3.1 inhibition, hyperexcitability, and seizures in acquired epilepsy. J Neurosci 42:5843–5859. 10.1523/JNEUROSCI.2475-21.2022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tobin VA, Douglas AJ, Leng G, Ludwig M (2011) The involvement of voltage-operated calcium channels in somato-dendritic oxytocin release. PLoS One 6:e25366. 10.1371/journal.pone.0025366 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tobin VA, Hurst G, Norrie L, Dal Rio FP, Bull PM, Ludwig M (2004) Thapsigargin-induced mobilization of dendritic dense-cored vesicles in rat supraoptic neurons. Eur J Neurosci 19:2909–2912. 10.1111/j.1460-9568.2004.03388.x [DOI] [PubMed] [Google Scholar]
- Tobin VA, Ludwig M (2007) The role of the actin cytoskeleton in oxytocin and vasopressin release from rat supraoptic nucleus neurons. J Physiol 582:1337–1348. 10.1113/jphysiol.2007.132639 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Torres GE, Arfken CL, Andrade R (1996) 5-Hydroxytryptamine4 receptors reduce afterhyperpolarization in hippocampus by inhibiting calcium-induced calcium release. Mol Pharmacol 50:1316–1322. [PubMed] [Google Scholar]
- Tzingounis AV, Kobayashi M, Takamatsu K, Nicoll RA (2007) The diffusible calcium sensor, hippocalcin, gates the calcium activation of the slow afterhyperpolarization in hippocampal pyramidal neurons. Neuron 53:487–493. 10.1016/j.neuron.2007.01.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ueta Y, et al. (2005) Transgenic expression of enhanced green fluorescent protein enables direct visualization for physiological studies of vasopressin neurons and isolated nerve terminals of the rat. Endocrinology 146:406–413. 10.1210/en.2004-0830 [DOI] [PubMed] [Google Scholar]
- van de Vrede Y, Fossier P, Baux G, Joels M, Chameau P (2007) Control of IsAHP in mouse hippocampus CA1 pyramidal neurons by RyR3-mediated calcium-induced calcium release. Pflugers Arch 455:297–308. 10.1007/s00424-007-0277-4 [DOI] [PubMed] [Google Scholar]
- Verkhratsky A, Shmigol A (1996) Calcium-induced calcium release in neurones. Cell Calcium 19:1–14. 10.1016/S0143-4160(96)90009-3 [DOI] [PubMed] [Google Scholar]
- Villalobos C, Andrade R (2010) Visinin-like neuronal calcium sensor proteins regulate the slow calcium-activated afterhyperpolarizing current in the rat cerebral cortex. J Neurosci 30:14361–14365. 10.1523/JNEUROSCI.3440-10.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Villalobos C, Shakkottai VG, Chandy KG, Michelhaugh SK, Andrade R (2004) SKCa channels mediate the medium but not the slow calcium-activated afterhyperpolarization in cortical neurons. J Neurosci 24:3537–3542. 10.1523/JNEUROSCI.0380-04.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vogalis F, Furness JB, Kunze WAA (2001) Afterhyperpolarization current in myenteric neurons of the Guinea pig duodenum. J Neurophysiol 85:1941–1951. 10.1152/jn.2001.85.5.1941 [DOI] [PubMed] [Google Scholar]
- Vogalis F, Harvey JR, Furness JB (2003) PKA-mediated inhibition of a novel K+ channel underlies the slow after-hyperpolarization in enteric AH neurons. J Physiol 548:801–814. 10.1113/jphysiol.2002.037325 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vogalis F, Harvey JR, Furness JB (2004) Suppression of a slow post-spike afterhyperpolarization by calcineurin inhibitors. Eur J Neurosci 19:2650–2658. 10.1111/j.0953-816X.2004.03369.x [DOI] [PubMed] [Google Scholar]
- Wang L, Chandaka GK, Foehring RC, Callaway JC, Armstrong WE (2018) Changes in potassium channel modulation may underlie afterhyperpolarization plasticity in oxytocin neurons during late pregnancy. J Neurophysiol 119:1745–1752. 10.1152/jn.00608.2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wingrove DE, Gunter TE (1986) Kinetics of mitochondrial calcium transport. II. A kinetic description of the sodium-dependent calcium efflux mechanism of liver mitochondria and inhibition by ruthenium red and by tetraphenylphosphonium. J Biol Chem 261:15166–15171. 10.1016/S0021-9258(18)66847-4 [DOI] [PubMed] [Google Scholar]
- Woods JJ, et al. (2019) A selective and cell-permeable mitochondrial calcium uniporter (MCU) inhibitor preserves mitochondrial bioenergetics after hypoxia/reoxygenation injury. ACS Cent Sci 5:153–166. 10.1021/acscentsci.8b00773 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Thapsigargin incubation inhibits VP neuron AHPs similarly in male and female rats. AHP parameters of 1 h TG incubation in male and female rats. Male control n = 8; male TG n = 17; female control n = 12; female TG n = 23. Statistics performed with a two-way ANOVA and Šídák's multiple comparisons test. P-values for the overall group comparison are provided in parentheses, for multiple comparisons refer to the figure. 0.1 s amplitude (F = 10.26, p < 0.0001), 1 s amplitude (F = 13.00, p < 0.0001), duration (F = 10.73, p < 0.0001), area (F = 12.20, p < 0.01), Ca2+ peak (F = 6.49, p < 0.01), Ca2+ area (F = 4.38, p = 0.01). Download Figure 1-1, TIF file (5.3MB, tif) .
Baseline Ca2+ measurements do not significantly differ after TPP+ or subsequent CCCP. Raw baseline fluorescence for Control, TPP+, and CCCP in both voltage clamp (left) and current clamp (right). In voltage clamp, we observed a significant difference (One-way repeated measures ANOVA, F = 4.80, p = 0.031) however post hoc analysis revealed no significant between group differences. In current clamp, we observed a significant difference (One-way repeated measures ANOVA, F = 4.11, p = 0.042) however post hoc analysis revealed no significant between group differences. Download Figure 6-1, TIF file (2.3MB, tif) .






