Skip to main content
The Journal of Neuroscience logoLink to The Journal of Neuroscience
. 2013 Apr 10;33(15):6476–6491. doi: 10.1523/JNEUROSCI.6384-11.2013

Separate Ca2+ Sources Are Buffered by Distinct Ca2+ Handling Systems in Aplysia Neuroendocrine Cells

Christopher J Groten 1, Jonathan T Rebane 1, Gunnar Blohm 1, Neil S Magoski 1,
PMCID: PMC6619096  PMID: 23575846

Abstract

Although the contribution of Ca2+ buffering systems can vary between neuronal types and cellular compartments, it is unknown whether distinct Ca2+ sources within a neuron have different buffers. As individual Ca2+ sources can have separate functions, we propose that each is handled by unique systems. Using Aplysia californica bag cell neurons, which initiate reproduction through an afterdischarge involving multiple Ca2+-dependent processes, we investigated the role of endoplasmic reticulum (ER) and mitochondrial sequestration, as well as extrusion via the plasma membrane Ca2+-ATPase (PMCA) and Na+/Ca2+ exchanger, to the clearance of voltage-gated Ca2+ influx, Ca2+-induced Ca2+-release (CICR), and store-operated Ca2+ influx. Cultured bag cell neurons were filled with the Ca2+ indicator, fura-PE3, to image Ca2+ under whole-cell voltage clamp. A 5 Hz, 1 min train of depolarizing voltage steps elicited voltage-gated Ca2+ influx followed by EGTA-sensitive CICR from the mitochondria. A compartment model of Ca2+ indicated the effect of EGTA on CICR was due to buffering of released mitochondrial Ca2+ rather than uptake competition. Removal of voltage-gated Ca2+ influx was dominated by the mitochondria and PMCA, with no contribution from the Na+/Ca2+ exchanger or sarcoplasmic/endoplasmic Ca2+-ATPase (SERCA). In contrast, CICR recovery was slowed by eliminating the Na+/Ca2+ exchanger and PMCA. Last, store-operated influx, evoked by ER depletion, was removed by the SERCA and depended on the mitochondrial membrane potential. Our results demonstrate that distinct buffering systems are dedicated to particular Ca2+ sources. In general, this may represent a means to differentially regulate Ca2+-dependent processes, and for Aplysia, influence how reproductive behavior is triggered.

Introduction

Intracellular Ca2+ transduces electrical signals into biochemical cascades that control vital processes, including gene expression, excitability, and secretion (Clapham, 1995). Free Ca2+ is determined by the equilibrium between Ca2+ sources and Ca2+ removal (Catterall and Few, 2008). In neurons, Ca2+ enters primarily through voltage-gated Ca2+ channels, although the endoplasmic reticulum (ER) and mitochondria provide additional Ca2+ reservoirs for release (Armstrong and Hille, 1998; Rizzuto and Pozzan, 2006). Ca2+ removal relies on plasma membrane extrusion, controlled by the Na+/Ca2+ exchanger and plasma membrane Ca2+-ATPase (PMCA), and sequestration, mediated by the mitochondrial uniporter and sarcoplasmic/endoplasmic reticulum Ca2+-ATPase (SERCA) (Kim et al., 2003, 2005; Rizzuto and Pozzan, 2006). The goal of this study is to examine the contribution of extrusion and sequestration to handling Ca2+ from different sources.

It is well established that the involvement of a given Ca2+ buffering system to the removal of a Ca2+ load varies between neuronal types and cytological compartments (Thayer and Miller, 1990; Fierro et al., 1998; Krizaj and Copenhagen, 1998; Morgans et al., 1998; Juhaszova et al., 2000; Holthoff et al., 2002; Kim et al., 2003, 2005). As the Ca2+-dependent activation of downstream targets relies on a specific concentration and/or pattern of Ca2+, this disparity likely reflects the different Ca2+ requirements for particular processes in a given neuron (e.g., motor vs sensory neuron) or compartment (e.g., dendrites vs axon terminal) (Caride et al., 2001; Berridge et al., 2003). Similar to different compartments, discrete Ca2+ sources can have unique roles in controlling neuronal function (Deisseroth et al., 1998; Berridge et al., 2000). Despite this, the relative contribution of given removal systems to the clearance of Ca2+ from distinct sources remains largely unexplored.

To address whether Ca2+ sources are handled uniquely, we used the bag cell neurons of the marine mollusk Aplysia californica. On stimulation, these neurons undergo ∼30 min of action potential firing, known as an afterdischarge, during which egg-laying hormone (ELH) is secreted into the blood stream to initiate reproduction (Kupfermann and Kandel, 1970; Arch, 1972; Pinsker and Dudek, 1977). As the afterdischarge progresses, voltage-gated Ca2+ influx, Ca2+-induced Ca2+-release (CICR), and store-operated Ca2+ influx are engaged to provide the Ca2+ that orchestrates a sustained increase in excitability and neuropeptide secretion through various Ca2+-dependent mechanisms (DeRiemer et al., 1984; Loechner et al., 1990; Wilson et al., 1996; Michel and Wayne, 2002; Kachoei et al., 2006; Hung and Magoski, 2007). We show that voltage-gated Ca2+ influx is primarily sequestered by the mitochondria, which subsequently releases the Ca2+ to ultimately be extruded across the plasma membrane. A second plasma membrane Ca2+ source, store-operated Ca2+ influx, is preferentially cleared by the SERCA. Recent evidence from cervical ganglion neurons indicates that the differential contribution of removal systems can control which Ca2+ source activates a given intracellular pathway (Wheeler et al., 2012). By analogy, the differential Ca2+ clearance we observe in bag cell neurons may facilitate the production of source-specific Ca2+ profiles. This could ensure coupling to specific Ca2+ signaling pathways in the presence of multiple, spatio-temporally overlapping Ca2+ signals.

Materials and Methods

Animals and cell culture

Adult Aplysia californica (a hermaphrodite) weighing 150–500 g were obtained from Marinus. Animals were housed in an ∼300 L aquarium containing continuously circulating, aerated artificial sea water (Instant Ocean, Aquarium Systems) at 14–16°C on a 12/12 h light/dark cycle and fed Romaine lettuce 5 times/week. For primary cultures of isolated bag cell neurons, animals were anesthetized by an injection of isotonic MgCl2 (∼50% body weight), the abdominal ganglion removed and incubated for 18 h at 22°C in neutral protease (13.33 mg/ml; 165859, Roche Diagnostics) dissolved in tissue culture artificial sea water (tcASW) (composition in mm: 460 NaCl, 10.4 KCl, 11 CaCl2, 55 MgCl2, 15 HEPES, 1 mg/ml glucose, 100 U/ml penicillin, and 0.1 mg/ml streptomycin, pH 7.8 with NaOH). The ganglion was then transferred to fresh tcASW and the bag cell neuron clusters were dissected from the surrounding connective tissue. Using a fire-polished Pasteur pipette and gentle trituration, neurons were dispersed onto 35 × 10 mm polystyrene tissue culture dishes (catalog #353001; Falcon, Becton Dickinson) filled with 2 ml of tcASW. Cultures were maintained in tcASW for 1–3 d in a 14°C incubator. Experiments were performed on neurons that were in vitro for at least 1 d. Salts were obtained from Fisher Scientific or Sigma-Aldrich.

Whole-cell, voltage-clamp recordings

Voltage-clamp recordings were made using an EPC-8 amplifier (HEKA Electronics) and the tight-seal, whole-cell method. Microelectrodes were pulled from 1.5 mm external, 1.2 mm internal diameter borosilicate glass capillaries (TW150F-4, World Precision Instruments) and had a resistance of 1–2.5 MΩ when filled with intracellular saline (see below). Pipette junction potentials were nulled, and subsequent to seal formation, pipette capacitive currents were cancelled. Following break-through, neuronal capacitance was also cancelled, and the series resistance (3–5 MΩ) compensated to 80% and monitored throughout the experiment. Current was filtered at 1 kHz with the EPC-8 Bessel filter and sampled at 2 kHz using a Digidata 1322A analog-to-digital converter (Molecular Devices), computer, and Clampex software (version 10.2, Molecular Devices). Voltage stimuli were delivered with Clampex.

Ca2+ currents were isolated using Ca2+-Cs+-tetraethylammonium (TEA) ASW, where the NaCl and KCl were replaced by TEA-Cl and CsCl, respectively, and the glucose and antibiotics were omitted (composition in mm: 460 TEA-Cl, 10.4 CsCl, 55 MgCl2, 11 CaCl2, 15 HEPES, pH 7.8 with CsOH). In some cases, the NaCl was not replaced by TEA to allow for Na+/Ca2+ exchanger activity. Whole-cell recordings used a Cs+-aspartate-based intracellular saline (composition in mm): 70 CsCl, 10 HEPES, 11 glucose, 10 glutathione, 5 ethyleneglycol bis (aminoethylether) tetraacetic acid (EGTA), 500 aspartic acid, 5 ATP (grade 2, disodium salt; A3377, Sigma-Aldrich), and 0.1 GTP (type 3, disodium salt; G8877, Sigma-Aldrich), pH 7.3 with CsOH. Certain experiments were performed with 0 mm EGTA in the internal saline. The majority of Ca2+-imaging (see below) was performed under whole-cell voltage clamp, during which the intracellular saline was supplemented with 1 mm fura-PE3 (0110; Teflabs) (Vorndran et al., 1995) to dye-fill neurons via passive dialysis. Cells were dialyzed for at least 20 min before experiments were performed.

Sharp-electrode current-clamp recording

For store-operated Ca2+ influx, current clamp was used to inject fura-PE3 into bag cell neurons using a PMI-100 pressure microinjector (Dagan). Neurons were filled with an optimal concentration of dye, similar to whole-cell conditions. Microelectrodes were pulled from 1.2 mm external, 0.9 mm internal diameter borosilicate glass capillaries (IB120F-4; World Precision Instruments) and had a resistance of 15–30 MΩ when the tip was filled with 10 mm fura-PE3 and backfilled with 3 m KCl. Recordings were made using an Axoclamp 2B amplifier (Molecular Devices) and the sharp-electrode, bridge-balanced method. All neurons used for imaging had resting potentials of −50 to −60 mV and action potentials that overshot 0 mV after depolarizing current injection (0.5–1 nA, from the amplifier). Store-operated influx recordings were performed in Ca2+-free ASW (composition as per tcASW but with added 0.5 mm EGTA and the glucose and antibiotics omitted) or Ca2+/Na+-free ASW [composition as per Ca2+-free ASW but with Na+ replaced with N-methyl-d-glucamine (NMDG)].

Ca2+ imaging

Imaging was performed using a TS100-F inverted microscope (Nikon) equipped with a Nikon Plan Fluor 10× [numerical aperture (NA) = 0.5], 20× (NA = 0.5), or 40× (NA = 0.6) objective. The light source was a 75 W Xenon arc lamp and a multiwavelength DeltaRAM V monochromatic illuminator (Photon Technology International) coupled to the microscope with a UV-grade liquid-light guide. Excitation wavelengths were 340 and 380 nm. Between acquisition episodes, the excitation illumination was blocked by a shutter, which along with the excitation wavelength, was controlled by a computer, a Photon Technology International computer interface, and EasyRatio Pro software (version 1.10, Photon Technology International). If image acquisition occurred at a frequency >0.2 Hz, the shutter remained open continuously. Emitted light passed through a 400 nm long-pass dichroic mirror and a 510/40 nm emission barrier filter before being detected by a Photometrics Cool SNAP HQ2 charge-coupled device camera. Camera gain was maximized and exposure time adjusted on a per cell basis. Exposure times during 340 and 380 excitation were fixed to the same value. Background was removed by setting a minimal threshold value of 300 arbitrary units of fluorescence. Fluorescence intensities were typically sampled at 0.5 Hz. For longer recordings, sampling was switched to 0.2 Hz, after any fast periods of Ca2+ dynamics. Fluorescence signals were acquired using regions of interest measured over neuronal somata, at approximately the midpoint of the vertical focal plane and one-half to three-quarters of the cell diameter, then averaged eight frames per acquisition. The ratio of the emission following 340 and 380 nm excitation (340/380) was taken to reflect free intracellular Ca2+ (Grynkiewicz et al., 1985), and saved for subsequent analysis. Image acquisition, emitted light sampling, and ratio calculations were performed using EasyRatio Pro.

Reagents and drug application

Solution exchanges were accomplished by manual perfusion using a calibrated transfer pipette to first exchange the bath (tissue culture dish) solution. In most cases where a drug was applied, a small volume (≤10 μl) of concentrated stock solution was mixed with a larger volume of saline (∼100 μl) that was initially removed from the bath, and this mixture was then pipetted back into the bath. Carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone (FCCP; 21857; Sigma-Aldrich), carboxyeosin (C-22803; Invitrogen), bafilomycin A (B1793, Sigma-Aldrich), and cyclopiazonic acid (CPA; C1530, Sigma-Aldrich or 239805, Calbiochem) all required dimethyl sulfoxide (DMSO; BP231, Fisher) as a vehicle. The maximal final concentration of DMSO was ≤0.5% (v/v) which, in control experiments as well as prior work from our laboratory, had no effect on membrane potential, various macroscopic or single-channel currents, resting intracellular Ca2+, or Ca2+ transients evoked by a train of action potentials (Kachoei et al., 2006; Lupinsky and Magoski, 2006; Hung and Magoski, 2007; Gardam et al., 2008; Geiger and Magoski, 2008; Tam et al., 2009, 2011; Hickey et al., 2010). Tetraphenylphosphonium chloride (TPP; 218790, Sigma-Aldrich) and lanthanum chloride (La3+; L-4131, Sigma-Aldrich) were prepared in water.

Analysis

Origin (version 7; OriginLab) was used to import and plot ImageMaster Pro files as line graphs. Analysis usually compared the steady-state value of the baseline 340/380 ratio with the ratio from regions that had reached a peak or new steady state. Averages of the baseline and peak regions were determined by eye or with adjacent-averaging. The rate of recovery of Ca2+ influx, after a stimulus train or store-operated Ca2+ entry and CICR, was quantified in different ways. To allow for comparison of data between our previous work on CICR (Geiger and Magoski, 2008), the rate of recovery from CICR was measured as the time required, after peak CICR, for the 340/380 ratio to return to 75% of the baseline ratio observed before the stimulus. The time at which the Ca2+ plateau first reached peak was considered time 0. Under circumstances where CICR was eliminated, time to 75% recovery was measured from the Ca2+ level at 1 min stimulation. This time was chosen because it reflects the typical point at which CICR responses peaked. Poststimulus area was used to quantify the magnitude and duration of CICR. Area was determined by integrating the region above the prestimulus baseline value from either 1 min post-train stimulus to 11 min or 11–21 min poststimulus. Again, measurements began at 1 min poststimulus to avoid including the initial recovery and capture peak CICR.

The degree of Ca2+ removal was also quantified by acquiring decay time constants and measuring the percentage recovery at 5 min poststimulus. Monoexponential decay functions were fit from the first point of decay to several minutes after complete recovery to baseline. The percentage recovery at 5 min was calculated by determining the degree of Ca2+ removed after the train stimulus or store-operated Ca2+ influx (340/380 peak–340/380 at 5 min post-peak) and dividing it by the peak rise during the response (340/380 peak–prestimulus baseline 340/380).

Summary data are presented as the mean ± SE). Statistics were performed using Instat (version 3.0; GraphPad Software). The Kolmogorov–Smirnov method was used to test datasets for normality. If the data were normal, Student's paired or unpaired t test (with the Welch correction as required) was used to test for differences between two means, whereas a standard one-way ANOVA with Dunnett's post hoc test was used to test for differences between multiple means. If the data were not normally distributed, a Mann–Whitney U test was used for two means, whereas a Kruskal–Wallis ANOVA with Dunn's post hoc test was used for multiple means. Fisher's exact test was used to test for differences in frequency between groups. A difference was considered significant if the two-tailed p value was <0.05.

For Figure 2, the rate of Ca2+ removal was determined for the post-train stimulus recovery period by deriving the slope of the Ca2+ decay ([Δ340/380]/Δt) at sequential time points using Microsoft Office Excel Plus 2010 (version 14). To prevent noise from influencing rate calculations, a fitted slope was measured starting at the initial decay point over 10 sequential time points ([Ca]n − [Ca]n+9/tntn+9) while incrementally shifting the start time (n+1) until the end of the decay phase. From this, a plot of Ca2+ decay rate versus 340/380 ratio was produced and fit with a second order polynomial function in Excel. Only traces that provided fits with r2 > 0.9 were used for further analysis. Multiple polynomial fits were used to make an average polynomial function describing the dataset.

Figure 2.

Figure 2.

Mitochondria remove voltage-gated Ca2+ influx and clear Ca2+ from repeated stimuli. A, Neurons voltage-clamped to −80 mV with 5 mm intracellular EGTA to allow for isolated measurement of voltage-gated Ca2+ influx and removal. A, Left, In DMSO, cytosolic Ca2+ transients evoked by the 5 Hz, 1 min train stimulus are followed by rapid recovery to baseline Ca2+. A, Right, Pretreatment with 20 μm FCCP, a protonophore that collapses the mitochondrial membrane potential and prevents Ca2+ sequestration, slows the recovery of Ca2+ following stimulation. A, Inset, The exponential decay time constant (τ) of the Ca2+ transient recovery phase is significantly larger in FCCP-treated neurons (unpaired Student's t test). B, Top, Relative Ca2+ clearance rate (R), calculated from the decay phase of Ca2+ transients shown in A, as a function of 340/380 ratio (rates normalized to the maximal value of the 340/380 range). Second-order polynomial fitted lines are plotted overtop of the data points. The difference between the control (Rtotal) and FCCP (RFCCP) fits produce the estimated mitochondrial uptake (Rmit), represented by the light dashed line. B, Bottom, Second-order polynomial fit lines for Rcontrol, RFCCP, and Rmit representing averaged removal rates from multiple neurons. Sample sizes are different from the decay time constants shown in A due to quality of fit criteria required for rate functions (see Materials and Methods). C, Ca2+ influx from a train stimulus loads mitochondria with Ca2+. C, Left, FCCP (20 μm) elevates Ca2+ in neurons under voltage clamp at −80 mV with 5 mm EGTA in the pipette and 100 μm TPP to prevent potential release of mitochondrial Ca2+. C, Right, FCCP-induced Ca2+ release after a large influx of Ca2+ from a 5 Hz, 1 min train stimulus is increased. D, Train stimulation, before FCCP application, significantly enhances the peak percentage change upon FCCP-induced Ca2+ liberation (unpaired Student's t test). E, Mitochondrial Ca2+ clearance is necessary for recovery from repeated stimuli. E, Left, After a train stimulus in FCCP, a second stimulus produces a Ca2+ load that is largely unremoved. E, Right, To replicate the slow Ca2+ recovery in FCCP, a control cell is given a train stimulus, but then subsequently held at −30 mV to allow for a small persistent Ca2+ influx. Ca2+ levels are quickly restored following a second train stimulus when the cell is clamped at −80 mV. F, The ratio between the first and second percentage recovery at 5 min is significantly larger in FCCP-treated neurons (unpaired Student's unpaired t test).

Model development

Equations describing mitochondrial Ca2+ dynamics were adapted from Colegrove et al. (2000b) to produce a compartment model of bag cell neuron Ca2+.

Plasma membrane Ca2+ flux.
graphic file with name zns01513-3642-m01.jpg
graphic file with name zns01513-3642-m02.jpg
graphic file with name zns01513-3642-m03.jpg

where Jinflux is the rate of Ca2+ influx across the plasma membrane, kinflux refers to the Ca2+ permeability of the membrane, and [Ca2+]i and [Ca2+]e are the intracellular and extracellular Ca2+ concentrations, respectively. To produce Ca2+ influx in the model, kinflux was transiently increased and then reduced manually. Jefflux is the rate of plasma membrane efflux, Vmax,efflux is the maximal rate of efflux, EC50, efflux is the Ca2+ concentration at which efflux is half-maximal, and nefflux is the Hill coefficient. Jpm is the net plasma membrane Ca2+ flux.

Mitochondrial Ca2+ dynamics.
graphic file with name zns01513-3642-m04.jpg
graphic file with name zns01513-3642-m05.jpg
graphic file with name zns01513-3642-m06.jpg
graphic file with name zns01513-3642-m07.jpg

where Juptake is the rate of mitochondrial Ca2+ sequestration, kmax, uptake is the mitochondrial uptake rate constant, EC50, uptake describes the Ca2+ concentration at which uptake is half-maximal, and nuptake is the Hill coefficient. The factor, δ([Ca2+]i), describes the inhibition of mitochondrial extrusion by cytosolic Ca2+. Kinhib is the Ca2+ concentration at which inhibition of Jrelease is half-maximal and ninhib describes the sensitivity of inhibition to cytosolic Ca2+. Vmax, release is the maximal rate of Ca2+ release from the mitochondria and EC50, release is the concentration of mitochondrial Ca2+ ([Ca2+]m) at which efflux rate is half of Vmax, release. Jmito is the net Ca2+ flux of the mitochondria.

Exogenous Ca2+ buffers.
graphic file with name zns01513-3642-m08.jpg

where JEGTA is the rate of free cytosolic Ca2+ removal by EGTA (Nowycky and Pinter, 1993), koff and kon are the reverse and forward reaction constants, respectively, [CaB] is the concentration of the Ca2+-EGTA complex, [Ca2+]i is the concentration of cytosolic Ca2+, and [B] is the concentration of free EGTA. Values for koff and kon (Table 1) were taken from Naraghi (1997), whereas [CaB] and [B] were calculated from the total EGTA concentration using MaxChelator (http://maxchelator.stanford.edu/CaEGTA-NIST.htm).

Table 1.

Parameter values used in compartment model of bag cell neuron Ca2+

Definition Model variable Standard value
Rate constant for PM Ca2+ influx kinflux 5 × 10−6 (s−1)
Extracellular Ca2+ concentration [Ca2+]e 11 mm
[Ca2+]i at half maximal rate of efflux EC50, efflux 378.8 nm
Hill coefficient for efflux nefflux 1.8
Maximal rate of efflux* Vmax, efflux 4.06 ± 0.34 nm/s (n = 15)
[Ca2+]i at half-maximal rate of mitochondrial uptake EC50, uptake 10 μm
Hill coefficient for mitochondrial Ca2+ uptake nuptake 2
Rate constant for mitochondrial Ca2+ uptake* kmax, uptake 10.3 ± 0.88 s−1 (n = 15)
[Ca2+]m at half-maximal rate of release EC50, release 307 nm
Maximal rate of mitochondrial Ca2+ release* Vmax, release 13.7 ± 4.0 nm/s (n = 15)
mitochondrial to cytosolic effective volume ratio γ 2
[Ca2+]i at half-maximal release inhibition EC50, inhib 500 nm
Hill coefficient for release inhibition ninhib 6
Dissociation constant of EGTA Kd, EGTA 180 nm
Forward rate constant of EGTA kon 2.7 ± 106 M−1 · s−1
Reverse rate constant of EGTA koff 0.5 s−1
Dissociation constant of fura Kd 760 nm
Endogenous Ca2+ binding ratio κe 60

*Indicates a free parameter estimated through data fitting (value ± SEM). All other parameters from the literature as indicated in text.

Ca2+-binding ratio.
graphic file with name zns01513-3642-m09.jpg

The variable, κ, represents the mean Ca2+-binding ratio over the Ca2+ range experienced during typical neuronal excitation (Neher and Augustine, 1992). [Bt] is the total buffer concentration, and Kd is the dissociation constant of the exogenous buffer. [Ca2+]i, rest and [Ca2+]i, peak are the free intracellular Ca2+ concentrations in the bag cell neurons at rest and during peak stimulus-induced influx, respectively.

Total Ca2+ removal rate.
graphic file with name zns01513-3642-m10.jpg

For model presentation, rates of change in free intracellular Ca2+ (d[Ca2+]F/dt) were converted to rates of total Ca2+ removal (d[Ca2+]T/dt). κB represents the average Ca2+-binding ratio for exogenous buffers (fura and EGTA) as calculated from Equation 9. Endogenous Ca2+-binding ratios (κS) were taken from estimates in Aplysia metacerebral neurons (Gabso et al., 1997).

Collective Ca2+ dynamics.
graphic file with name zns01513-3642-m11.jpg
graphic file with name zns01513-3642-m12.jpg

where d[Ca2+]i/dt is the rate of change in cytosolic Ca2+, d[Ca2+]m/dt is the rate of change of mitochondrial Ca2+, and γ is the ratio of effective mitochondrial and cytoplasmic volumes. The γ value used was taken from estimates in bullfrog sympathetic neurons (Colegrove et al., 2000b). For the estimates of model parameters, 340/380 ratios were converted to values of free intracellular Ca2+ based on Ca2+-sensitive electrode recordings and fura Kd measurements in Aplysia (Fisher et al., 1994; Gabso et al., 1997; Michel and Wayne, 2002). To fit individual traces, the Vmax, efflux, kmax, uptake, and Vmax, release were left as free variables, whereas the constants (EC50 values) describing the Ca2+ sensitivity of plasma membrane extrusion and mitochondrial Ca2+ release were set to parameters established in bullfrog sympathetic neurons (Colegrove et al., 2000a,b). The components describing mitochondrial Ca2+ uptake (EC50, uptake and nuptake) were based on measurements in isolated mitochondria and used to determine kmax,uptake (Gunter and Pfeiffer, 1990; Gunter and Gunter, 1994; Colegrove et al., 2000b). Free parameters were then optimized to fit individual experimental records. Perhaps because of the time required for Ca2+ diffusion from the plasma membrane to the bulk of the cytosolic mitochondria, we found that proper fitting of bag cell neuron CICR often required a delayed onset of mitochondrial Ca2+ uptake. To account for this, our model implemented a time delay between the initial Ca2+ influx and the onset of mitochondrial buffering. The parameter estimates obtained from fitting were then collected from multiple neurons and averaged to obtain the values presented in Table 1. These values are shown as free, not total, intracellular Ca2+ to allow for comparison between similar models. Differential equations were solved numerically using Euler's method written in MATLAB with a time-step of 70 ms to produce graphical outputs of total cytosolic and mitochondrial Ca2+ over time.

Results

Mimicking the fast phase of the afterdischarge evokes distinct Ca2+ dynamics in bag cell neurons

A brief input to the bag cell neurons initiates the afterdischarge: a prolonged period of action potential firing consisting of a fast phase of ∼5 Hz for ∼1 min, which progresses into a slow phase of ∼1 Hz for ∼30 min (Kaczmarek et al., 1982; Fisher et al., 1994). To examine Ca2+ dynamics in response to a fast phase-like stimulus, a 1 min, 5 Hz train of 75 ms depolarizing steps to 0 mV was applied to fura-PE3-loaded, cultured bag cell neurons from a holding potential of −80 mV under whole-cell voltage clamp. Unless stated otherwise, all neurons were recorded using a Cs+-containing and TEA-containing external solution (to replace K+ and Na+, respectively) and a Cs+-containing internal pipette solution (to replace intracellular K+; see Materials and Methods for details).

Application of the train stimulus produced a large, transient rise in intracellular Ca2+ due to the activation of voltage-gated Ca2+channels, followed by an exponential decline, with recovery to baseline in 5–10 min (n = 8) (Fig. 1A, left). This response was measured with our Cs+-based intracellular saline containing 5 mm Ca2+ chelator, EGTA. As this buffer alters free intracellular Ca2+, we sought to apply the same stimulus when EGTA was omitted from the pipette solution. With 0 mm intracellular EGTA, excitation again resulted in a large Ca2+ transient; however, a prolonged Ca2+ plateau, often marked by a delayed peak, now followed the initial recovery (n = 6) (Fig. 1A, right). This Ca2+ plateau long outlasted the duration of the stimulus, and was followed by a slow return to baseline within 10–20 min. Similar sequences of changes to intracellular Ca2+ have been described as CICR in dorsal root ganglion neurons, bullfrog sympathetic neurons, adrenal chromaffin cells, and Aplysia neuron R15 (Gorman and Thomas, 1980; Friel and Tsien, 1994; Herrington et al., 1996; Colegrove et al., 2000a).

Figure 1.

Figure 1.

A train of depolarizing stimuli induces a secondary Ca2+ rise sensitive to the Ca2+ chelator, EGTA. A, Simultaneous measurement of free intracellular Ca2+ and membrane current in cultured bag cell neurons using 340/380 fura PE3 fluorescence and whole-cell voltage clamp at −80 mV. A, Inset, A phase contrast image shows the recording pipette, bag cell neuron soma, and its neuritic processes. The bottom image shows the same neuron loaded with fura and the somatic region of interest (ROI) used for data collection. Scale bar applies to both images. A, Top left, Ca2+ influx indicated by a change in intensity of the 340/380 fluorescence ratio following a 1 min, 5 Hz train of 75 ms steps from −80 to 0 mV. With 5 mm intracellular EGTA, stimulation causes a large rise in Ca2+ followed by a rapid recovery to the prestimulus baseline. Top right, In the absence of intracellular EGTA, there is a prolonged Ca2+ plateau subsequent to the initial influx that greatly outlasts the stimulus duration, indicative of CICR. A, Bottom, Traces depict 300 overlaid Ca2+ currents from each pulse to 0 mV of the 1 min train stimulus in either 5 or 0 mm intracellular EGTA. The shifting band of traces is due to use-dependent inactivation of Ca2+ currents during the train stimulus. Unless stated otherwise, all cells were recorded in a Cs+-external and TEA-external (to replace K+ and Na+, respectively) and a Cs+-based intracellular solution (to replace intracellular K+). B, Left, The percentage change in 340/380 from baseline to the peak response during the train stimulus is significantly larger in 5 mm EGTA versus 0 mm EGTA (unpaired Student's t test). For this and subsequent bar graphs, data represents the mean ± SE, and the n-value is indicated within the bars. B, Middle and right, Zero mm EGTA significantly increases the total area measured from 1 min after stimulation to 11 min post-train stimulus (Mann–Whitney U test) and the time to reach 75% recovery to baseline Ca2+ from the peak of the plateau (Mann–Whitney U test).

Removing intracellular EGTA significantly reduced the peak percentage change in intracellular Ca2+ during the train stimulus compared with 5 mm EGTA (Fig. 1B, left). This is likely due to a facilitation of Ca2+-dependent inactivation of voltage-gated Ca2+ channels and an increase in resting Ca2+ levels (5 mm EGTA resting 340/380: 0.18 ± 0.004, n = 8; 0 mm EGTA resting 340/380: 0.26 ± 0.01, n = 6; p < 0.003, unpaired Mann–Whitney U test). The area from 1 to 11 min poststimulus train (10 min total) was used to quantify the magnitude and duration of the Ca2+ plateau (see Materials and Methods for details). The presence of the Ca2+ plateau in 0 mm EGTA was reflected by a significant increase in poststimulus train area from 1 to 11 min (Fig. 1B, middle) and the time to 75% recovery from peak post-train stimulus Ca2+ (Fig. 1B, right). Thus, mimicking an endogenous firing pattern evoked distinct rapid and slow periods of cytosolic Ca2+ dynamics in the bag cell neurons. As the transduction of a Ca2+ signal to activate unique biochemical pathways relies on the temporal and spatial properties of intracellular Ca2+, we sought to dissect the Ca2+ sources and removal processes that contributed to these Ca2+ responses.

Voltage-gated Ca2+ influx from a train stimulus is cleared by mitochondrial uptake

Mitochondria are an essential Ca2+ removal system in many neurons and neuroendocrine cells, particularly when Ca2+ concentrations are substantially higher than at rest (>500 nm) (Herrington et al., 1996). Prior work by our lab indicated a role for mitochondrial Ca2+ uptake after a train of action potentials (Geiger and Magoski, 2008). To test whether the mitochondria are key to removal of voltage-gated Ca2+ influx, the ability of mitochondria to clear Ca2+ was eliminated using FCCP. This protonophore collapses the mitochondrial membrane potential, releases stored Ca2+, and prevents subsequent Ca2+ uptake into the organelle (Heytler and Prichard, 1962; Babcock et al., 1997). EGTA (5 mm) was included in the pipette solution to eliminate the Ca2+ plateau and allow for isolated measurement of voltage-gated Ca2+ influx and removal. Post-train stimulus recoveries were well fit with monoexponential decay functions, from which time constants were derived to quantify changes in Ca2+ clearance. Percentage recovery at 5 min was measured to quantitate the degree of Ca2+ recovery. Any sample size discrepancy between percentage recovery and decay constants for the same dataset was because of the exclusion of poor exponential fits.

Initial observations showed that some neurons treated with FCCP had a reduced peak rise in Ca2+ during the train stimulus. This may be because of an increase in use-dependent inactivation of Ca2+ currents in the absence of mitochondrial Ca2+ clearance. To prevent this from impacting quantification of Ca2+ removal, additional FCCP-treated neurons were stimulated with 5 Hz, 1 min train of 175 ms pulses to enhance Ca2+ influx and match the peak levels seen in controls. Cells that presented peak Ca2+ amplitudes comparable to control were used in measuring the percentage recovery. As such, peak Ca2+ influx was not significantly different between DMSO-treated and selected FCCP-treated neurons (DMSO peak % Δ 340/380: 135.4 ± 8.8, n = 10; FCCP peak % Δ 340/380: 162.3 ± 14.5, n = 12; p > 0.05, unpaired Student's t test).

Compared with DMSO-treated cells, neurons stimulated after a 30 min exposure to 20 μm FCCP presented a slower Ca2+ recovery (Fig. 2A), as indicated by a significantly larger decay time constant (Fig. 2A, inset). These findings are consistent with the effects of FCCP in other systems, and indicate a role for mitochondrial Ca2+ clearance (Thayer and Miller, 1990; Friel and Tsien, 1994; Werth and Thayer, 1994). From the decay of these Ca2+ transients, we determined the relative rate of apparent mitochondrial uptake (Rmit) by subtracting the rate of Ca2+ removal in FCCP conditions (RFCCP) from the total Ca2+ removal rate (Rtotal) at corresponding 340/380 ratio values (see Materials and Methods for details). Figure 2B (top) displays the relative cytosolic Ca2+ removal rate, normalized to peak rate, against the 340/380 ratio for the representative traces in Figure 2A, along with fitted polynomial functions. Fits from multiple neurons were used to produce averaged RFCCP, Rtotal, and Rmit values (Fig. 2B, bottom). This suggests that mitochondrial uptake occurred over a wide range of Ca2+ levels, both at rest and at peak values during stimulation, with a corresponding increase in removal rate. In contrast, the nonmitochondrial Ca2+ buffer, represented as RFCCP, had a shallower slope over the same range, indicating the presence of a relatively slow removal mechanism.

Central to the notion that the mitochondria buffer voltage-gated Ca2+ influx, is that there is an increase in mitochondrial Ca2+ after stimulation. To examine this, FCCP was used to liberate mitochondrial Ca2+, with or without a prior train stimulus. Assuming mitochondrial involvement, 100 μm tetraphenylphosphonium (TPP), a blocker of mitochondrial Ca2+ exchange in bag cell neurons (Karadjov et al., 1986; Geiger and Magoski, 2008), was applied to both DMSO-treated and FCCP-treated cells, to ensure no mitochondrial Ca2+ release followed the excitation. As per prior work suggesting the mitochondria of cultured bag cell neurons contain Ca2+ at rest (Jonas et al., 1997; Gardam et al., 2008; Geiger and Magoski, 2008), bath application of 20 μm FCCP increased cytosolic Ca2+ within 5 min (n = 6) (Fig. 2C, left). If a train stimulus was delivered before FCCP (n = 6), the Ca2+ release signal was significantly increased by ∼40%, consistent with voltage-gated Ca2+ influx enhancing the amount of Ca2+ stored in the mitochondria (Fig. 2C, right, D).

As mitochondria appear to be essential for Ca2+ removal after depolarization, we attempted to saturate Ca2+ clearance with and without active mitochondria. After a 5 Hz, 1 min train stimulus in FCCP (n = 5), Ca2+ often decayed slowly to a higher level than the prestimulus baseline (Fig. 2E). Application of a second train stimulus from this new baseline elevated Ca2+ to similar levels as the first train stimulus; however, the subsequent recovery was severely hindered (Fig. 2E, left). To control for differences between FCCP and control in Ca2+ after the first train stimulus, DMSO-treated neurons were held at a potential ranging from −20 to −30 mV immediately after the first train stimulus (n = 6). Because some neurons had smaller Ca2+ currents, a range of voltages was used to ensure that all Ca2+ plateaus were of comparable size. Voltage-clamping at the depolarized potential produced an elevated Ca2+ baseline similar to that seen in FCCP. After the second train stimulus, cells were again held at −80 mV, but unlike in FCCP, DMSO-treated cells recovered rapidly to the Ca2+ levels as seen at the start of the experiment (Fig. 2E, right). The ratio between the percentage recovery at 5 min after the first and second train stimulus was used to quantify the degree of buffer saturation in each condition. The percentage recovery ratio was significantly reduced in FCCP-treated compared with DMSO-treated neurons (Fig. 2F).

Contribution of the plasma membrane Ca2+ ATPase to the removal of voltage-gated Ca2+ influx

Although our results strongly suggest that mitochondria are the predominant buffer for voltage-gated Ca2+ influx, a residual contribution from other systems must exist, given the slow recovery even in the presence of FCCP. Most cells use the high-affinity, low-capacity PMCA to extrude Ca2+ across the plasma membrane (Sanchez-Armass and Blaustein, 1987; Blaustein and Lederer, 1999; Jeon et al., 2003; Tidow et al., 2012). The role of this pump in removing voltage-gated Ca2+ influx was tested with 2 mm extracellular La3+, a common PMCA inhibitor (Carafoli, 1991; Herrington et al., 1996; Zenisek and Matthews, 2000). Addition of La3+ ∼1 s subsequent to the end of the 5 Hz, 1 min train stimulus (n = 16) slowed the recovery to prestimulus baseline (Fig. 3A, left) compared with control neurons treated with water (n = 17). This manifested as a significantly smaller percentage recovery at 5 min poststimulus in La3+-treated neurons (Fig. 3C). Consistent with a minor role of the PMCA in removing Ca2+, La3+ appeared to have a smaller effect on percentage recovery than FCCP (Fig. 3C). Because La3+ was bath applied on the last pulse of the train stimulus, to avoid blocking Ca2+ currents, the onset of La3+ action was delayed for several seconds thereafter. Thus, post-train stimulus decay time constants were not determined for this experiment, as they would not have been an accurate reflection of PMCA inhibition.

Figure 3.

Figure 3.

Removal of voltage-gated Ca2+ influx is slowed by inhibition of the PMCA, but is not influenced by acidic stores, ER, or the Na+/Ca2+-exchanger. A, Left, Blocking PMCA function by the addition of 2 mm extracellular La3+ on the last pulse of the 1 min, 5 Hz train stimulus, hinders Ca2+ removal (dark trace), whereas H2O (control) (light trace) applied in the same manner has no effect. A, Middle left, In contrast, the inclusion of Na+ (dark trace) rather than TEA (light trace) in the extracellular saline, to permit Na+/Ca2+ exchanger activity, has no effect. Similarly, treatment with 20 μm CPA (middle right), to inhibit SERCA, or 100 nm bafilomycin A (baf) (right), to prevent uptake by acidic stores, does not alter Ca2+ removal after a train stimulus. B, Neurons treated with 20 μm FCCP present significantly larger exponential decay time constants (τ) relative to controls (reproduced from Fig. 2A), whereas cells in the presence of extracellular Na+, exposed to CPA, or baf do not have significantly different τ values (all comparisons using unpaired Student's t test). C, FCCP (Mann–Whitney U test) and La3+ significantly (Welch corrected unpaired Student's t test) reduce the percentage recovery at 5 min post-train stimulus whereas extracellular Na+, CPA, and baf have no effect (all comparisons using unpaired Student's t test).

The plasma membrane Na+/Ca2+ exchanger trades extracellular Na+ for intracellular Ca2+ (Blaustein and Lederer, 1999; Kim et al., 2003). In the bag cell neurons, Na+/Ca2+ exchanger activity can be eliminated by replacing extracellular Na+ with TEA (Knox et al., 1996). Our standard recording conditions used extracellular TEA, rather than Na+; thus, we tested the effect of supplementing extracellular Na+ in lieu of TEA on the rate of voltage-gated Ca2+ removal (Fig. 3A). Compared with TEA (n = 6), adding extracellular Na+ (n = 7) did not significantly alter the magnitude of the Ca2+ rise during stimulation (TEA peak % Δ: 174.9 ± 18.8, n = 6; Na+ peak % Δ: 150.5 ± 22.3, n = 7; p > 0.05, unpaired Student's t test), poststimulus Ca2+ decay time constant (Fig. 3B) or the percentage recovery at 5 min after peak Ca2+ (Fig. 3C).

In addition to the mitochondria, the ER is the other primary intracellular Ca2+ store in neurons (Berridge et al., 2000), and has been found to remove voltage-gated Ca2+ influx in neurons and neuroendocrine cells (Fierro et al., 1998; Kim et al., 2003). To test this in bag cell neurons, 20 μm CPA, a SERCA inhibitor found to be effective in bag cell neurons (Seidler et al., 1989; Kachoei et al., 2006; Gardam et al., 2008; Geiger and Magoski, 2008), was applied 30 min before stimulation. Post-train stimulus Ca2+ kinetics were not affected by the presence of CPA (n = 6) versus control (n = 7) (Fig. 3A). CPA did not alter the peak rise in Ca2+ during stimulation (DMSO peak % Δ: 174.3 ± 20.4, n = 7; CPA peak % Δ: 164.8 ± 26.2, n = 6; p > 0.05, unpaired Student's t test), the poststimulus Ca2+ decay time constant (Fig. 3B), or the percentage recovery to baseline following stimulation (Fig. 3C).

Finally, FCCP also collapses other stores with proton gradients, including lysosome, endosomes, and secretory vesicles (Goncalves et al., 1999; Christensen et al., 2002). The contribution of these stores to the removal of voltage-gated Ca2+ influx was tested by treating cells with bafilomycin A, a H+-ATPase inhibitor that prevents the sequestration of Ca2+ by acidic stores (Bowman et al., 1988; Goncalves et al., 1999). Our earlier work demonstrated that bafilomycin A causes a slow, steady increase in bag cell neuron cytosolic Ca2+, distinct from the response to other liberating agents (Kachoei et al., 2006; Hickey et al., 2010). Pretreatment with 100 nm bafilomycin A (n = 6) did not alter the post-train stimulus Ca2+ removal compared with DMSO-treated neurons (n = 6) (Fig. 3A). Bafilomycin did not change the post-stimulus decay time constant (Fig. 3B), the percentage recovery at 5 min after peak Ca2+ (Fig. 3C) or the peak Ca2+ rise (DMSO peak % Δ: 235.58 ± 18.3, n = 6; bafilomycin A peak % Δ: 233.0 ± 23.7, n = 6; p > 0.05, unpaired Student's t test). Thus, the effect of FCCP on Ca2+ removal was due its action on mitochondrial function.

The PMCA, but not the SERCA or Na+/Ca2+ exchanger, clear somatic Ca2+ in the absence of mitochondrial function

The inhibition of a dominant Ca2+ clearance system can unveil the activity of other, formerly uninvolved removal mechanisms (Zenisek and Matthews, 2000; Kim et al., 2005). This presumably reflects a compensatory property that ensures normal Ca2+ homeostasis. We examined this possibility in bag cell neurons by exploring the contribution of nonmitochondrial clearance mechanisms in the presence of FCCP. Addition of the PMCA inhibitor, La3+, to the extracellular solution on the last pulse of the 5 Hz, 1 min train stimulus (n = 7), significantly blunted the recovery from peak compared with FCCP alone (n = 5). In the presence of FCCP, once full PMCA inhibition manifested, little to no recovery occurred and Ca2+ remained at a much higher plateau than in FCCP alone (Fig. 4A, left). The percentage recovery at 5 min was significantly reduced under these conditions (Fig. 4C). These data indicates that in the absence of both mitochondrial and PMCA function, voltage-gated Ca2+ removal is largely occluded. In contrast, exchanging extracellular Na+ (n = 7) for TEA (n = 8), or pretreatment with CPA (control, n = 8; CPA, n = 8) remained ineffective at influencing post-train stimulus Ca2+ removal in the absence of mitochondrial function (Fig. 4A, middle, right). Poststimulus decay time constants (Fig. 4B) and percentage recoveries at 5 min (Fig. 4C) were unchanged by including extracellular Na+ or pretreatment with CPA.

Figure 4.

Figure 4.

In the absence of mitochondrial function, PMCA, but neither SERCA nor the Na+/Ca2+-exchanger, influence the removal of voltage-gated Ca2+ influx. A, Left, Post-stimulus addition of the PMCA inhibitor, La3+ (2 mm) (dark trace), in the presence of 20 μm FCCP, strongly attenuates the subsequent recovery relative to FCCP alone (light trace). The inclusion of extracellular Na+ (middle) or pretreatment with 20 μm CPA (right) does not influence Ca2+ removal in the presence of FCCP. B, For neurons pretreated with FCCP, the mean post-train stimulus recovery τ values are not significantly altered by the inclusion of extracellular Na+ or exposure to CPA (unpaired Student's t test and Mann–Whitney U test, respectively). C, Summary of mean percentage recovery at 5 min post-train stimulus. In the presence of FCCP, La3+ significantly increases the percentage recovery at 5 min, whereas the application CPA (Mann–Whitney U test) or the inclusion of extracellular Na+ remains ineffective (all other comparisons using unpaired Student's t tests).

The EGTA-sensitive Ca2+ plateau is caused by mitochondrial Ca2+ release

In many neurons, it is common for brief periods of action potential firing to evoke sustained Ca2+ release from mitochondria or the ER with similar characteristics as the EGTA-sensitive Ca2+ plateau shown in Figure 1 (Gorman and Thomas, 1980; Smith et al., 1983; Neering and McBurney, 1984; Tang and Zucker, 1997; Lee et al., 2007). Prior research from our lab has found that prolonged stimulation of bag cell neurons under sharp electrode recording elicited a CICR plateau that was sensitive to FCCP as well as TPP, an inhibitor of mitochondrial Ca2+ exchangers (Geiger and Magoski, 2008). To determine whether the EGTA-sensitive Ca2+ plateau we observed under whole-cell conditions was also due to mitochondrial Ca2+ release, 100 μm TPP was applied to cells 30 min before stimulation. In the presence of TPP (n = 8), the post-train stimulus response transformed from a slow, large Ca2+ plateau under control conditions (n = 6) to a rapid exponential recovery indistinguishable from that seen in 5 mm intracellular EGTA (compare Figs. 5A, 1A). This was apparent from the significantly reduced post-train stimulus area from 1 to 11 min (Fig. 5C) and time to 75% recovery from peak Ca2+ (Fig. 5D). TPP had this effect without altering the peak rise in Ca2+ during stimulation (control peak % Δ: 127.5 ± 7.4, n = 6; TPP peak % Δ: 127.6 ± 11.5, n = 8; p > 0.05 unpaired Student's t test).

Figure 5.

Figure 5.

The EGTA-sensitive Ca2+ plateau is caused by mitochondrial Ca2+ release. A, A post-stimulus Ca2+ plateau is evoked by a 5 Hz, 1 min train of depolarizing steps under voltage clamp. Intracellular EGTA is absent (0 mm) to allow for CICR. The prolonged Ca2+ plateau in control conditions (light trace) is substantially smaller in 100 μm the mitochondrial Ca2+ exchange blocker, TPP (dark trace). B, Treatment with CPA, to inhibit SERCA and prevent ER Ca2+ release, does not affect the post-train stimulus Ca2+ plateau (dark trace) when compared with DMSO-treated neurons (light trace). C, D, TPP but not CPA, significantly reduces the 1–11 min area relative to controls (unpaired Student's t test in both cases). TPP also significantly reduces the time to reach 75% recovery of baseline Ca2+ from peak CICR (Welch corrected, unpaired Student's t test). In contrast, CPA has no significant effect on the recovery time (unpaired Student's t test). Sample sizes for the time to 75% recovery in CPA are smaller than those used for post-stimulus area measurements as some neurons did not achieve 75% recovery.

To determine whether CICR from the ER contributed to the plateau, CPA was used to deplete the ER of Ca2+ before stimulation. Cells treated with 20 μm CPA (n = 7) presented a similar Ca2+ plateau magnitude as in control conditions (n = 8) and recovered to prestimulus baseline with a comparable time course (Fig. 5B). CPA did not significantly alter post-train stimulus area from 1 to 11 min (Fig. 5C) and did not affect the time to 75% recovery from peak Ca2+ (Fig. 5D). These results indicate that the EGTA-sensitive Ca2+ plateau is mitochondrial, but not ER, Ca2+ release, and that removal of CICR is independent of the SERCA.

CICR is removed by plasma membrane extrusion via Na+/Ca2+ exchange and the PMCA

As the Ca2+ plateau from mitochondrial release had a different magnitude and kinetics than voltage-gated Ca2+ influx during the train stimulus, it is possible that the handling mechanisms responsible for CICR are different from those for voltage-gated Ca2+. Full recovery from CICR typically required >10 min; therefore, we presumed that a relatively slow clearance system was responsible for its removal. To examine this, we first tested for the contribution of the Na+/Ca2+ exchanger by substituting extracellular Na+ for TEA. In contrast to its ineffectiveness in removing voltage-gated Ca2+ influx, extracellular Na+ significantly reduced the time to 75% recovery from peak CICR to baseline (Fig. 6A, left) compared with TEA. However, the post-train stimulus area from 1 to 11 min was not significantly different between Na+ and TEA conditions (Fig. 6B). This occurred because the first stage of CICR is marked by the rising Ca2+ plateau, which was not significantly different in magnitude between TEA and Na+ conditions (TEA external, peak % Δ: 86.3 ± 2.0, n = 7; Na+ external, peak % Δ: 79.6 ± 5.0, n = 12; p > 0.05, Mann–Whitney U test). Nevertheless, the area from 11 to 21 min post-train stimulus, where the recovery from peak is most prominent, was significantly smaller in the presence of extracellular Na+ (Fig. 6B). Sample size between time to 75% recovery and post-train stimulus area for Na+-conditions was different because one cell did not recover to 75% by the end of the recording.

Figure 6.

Figure 6.

Mitochondrial Ca2+ release is removed by the Na+/Ca2+ exchanger and the PMCA. A, Left, In the absence of extracellular Na+ (TEA, light trace), the mitochondrial Ca2+ plateau evoked by a 1 min, 5 Hz train stimulus from −80 mV is prolonged compared with cells bathed in extracellular Na+ (dark trace). A, Inset, The presence of extracellular Na+ significantly accelerated the time to 75% recovery for CICR (Mann–Whitney U test). A, Middle, In Na+-free extracellular saline (TEA), the application of La3+ (dark trace) during the onset of CICR occludes recovery. In these experiments, cells are voltage-clamped at −80 mV following the stimulus. A, Right, As a control, a 1 min 5 Hz train stimulus from −80 mV, followed by a prolonged step depolarization evokes a Ca2+ plateau similar to CICR (at break, a portion of the Ca2+ influx is omitted for clarity to emphasize the Ca2+ plateau). The recovery of this Ca2+-influx plateau is insensitive to the replacement of extracellular Na+ (dark trace) for TEA (light trace). For this experiment, neurons are pretreated with 100 μm TPP to prevent any influence of mitochondrial Ca2+ release and are recorded with 0 mm intracellular EGTA. B, The post-train stimulus area from 1 to 11 min during mitochondrial Ca2+ release is not significantly different between Na+ and TEA conditions (unpaired Student's t test). However, the area from 11 to 21 min is significantly larger in TEA-containing extracellular saline (Mann–Whitney U test). C, In the absence of Na+/Ca2+ exchanger activity, the application of 2 mm La3+ during CICR does not alter the post-train stimulus area from 1 to 11 min (Mann–Whitney U test), but significantly increases the area from 11 to 21 min (Welch corrected Student's t test, respectively). D, Area summary data for the Ca2+ plateau evoked by post-train stimulus depolarization at (ranging from −10 to −20 mV). Replacing extracellular Na+ with TEA does not significantly alter the post-train stimulus area from time 1–11 min or 11–21 (Student's t test for both).

Even in the presence of TEA, when the Na+/Ca2+ exchanger was inhibited, CICR recovery still occurred, albeit at a slower rate, indicating involvement of another removal system. To determine whether the PMCA was responsible, 2 mm La3+ was applied at the peak of CICR, ∼1 min post-train stimulus, while in the presence of extracellular TEA. PMCA inhibition by La3+ halted CICR recovery (n = 7) whereas in control cells (n = 8) Ca2+ still returned to baseline (Fig. 6A, middle). As with the Na+-replacement experiment, the post-train stimulus area from 1 to 11 min was not significantly different in La3+-treated neurons, whereas the area from 11 to 21 min post-train stimulus was significantly enhanced (Fig. 6C).

The activation of Ca2+-dependent processes is highly sensitive to magnitude, duration, and frequency of cytosolic Ca2+ change (Clapham, 1995; Berridge et al., 2003). Therefore, the source specific involvement of the Na+/Ca2+ exchanger may be related to the difference between the rapid, large voltage-gated Ca2+ influx, and the slow, moderately sized CICR from the mitochondria. To test this, we examined whether a voltage-gated Ca2+ influx plateau, that was similar in amplitude and kinetics to CICR, was sensitive to extracellular Na+. Neurons were stimulated with a 5 Hz, 1 min train stimulus from −80 mV; however, unlike control conditions, cells were immediately voltage-clamped at a potential ranging from −10 to −20 mV for the remainder of the recording. Holding at a depolarized potential produced a steady Ca2+ influx comparable to that elicited during CICR. As with rapid, train stimulus-induced voltage-gated Ca2+ influx, the inclusion of extracellular Na+ (n = 5) had no apparent effect on the recovery of the slow, persistent voltage-gated Ca2+ influx plateau compared with controls (n = 5) (Fig. 6A, right). The post-train stimulus areas from 1 to 11 min, and 11–21 min were not significantly different between TEA and Na+ external conditions (Fig. 6D).

A model of bag cell neuron Ca2+ dynamics recapitulates EGTA-sensitive CICR

The data concerning bag cell neuron Ca2+ removal and release after prolonged stimulation provided the information necessary to create a model of Ca2+ dynamics. Previous work from bullfrog sympathetic neurons demonstrated that patterns of Ca2+ influx and release can be accounted for by a 3 component model consisting of extracellular, cytosolic, and mitochondrial compartments (Friel and Tsien, 1994; Colegrove et al., 2000b). Therefore, the parameter framework from these models was used in this study. Our three-compartment Ca2+ model included an extracellular Ca2+ influx source (Jinflux), uptake (Juptake) and release (Jrelease) by a mitochondrial store, and plasma membrane Ca2+ efflux (Jefflux) (Fig. 7A, inset). For simplification, extrusion by the PMCA and Na+/Ca2+ exchanger was represented by a single component. Estimates of these rates in the bag cell neurons were derived by fitting raw measurements of free intracellular Ca2+ (Fig. 7A) (see Materials and Methods). Values derived from individual traces were collected and averaged to produce the model parameters used for the subsequent graphs (Fig. 7B–D). To account for the buffering of Ca2+ by EGTA, fura, and endogenous Ca2+-binding proteins, rates of free Ca2+ removal and extrusion were converted to rates of total Ca2+ removal. Thus, the model output is presented as concentrations of total cytosolic Ca2+ ([Ca2+]T) (Fig. 7B–D) (see Materials and Methods).

Figure 7.

Figure 7.

A model of bag cell neuron Ca2+ dynamics demonstrates the sensitivity of mitochondrial Ca2+ release to the Ca2+ chelating agent, EGTA. A, The inset illustrates the components involved in the 3 compartment model of Ca2+. Ca2+ influx is represented by Jinflux, cytosolic Ca2+ removal is mediated by mitochondrial uptake (Juptake) or plasma membrane efflux (Jefflux), and mitochondrial Ca2+ release is denoted as Jrelease. Collectively, these variables determine the cytosolic and mitochondrial Ca2+concentrations (see Materials and Methods). Raw data from a bag cell neuron (light trace) presenting stimulus-induced Ca2+ influx, CICR, and recovery is fit by the model (dark trace) to estimate values for free intracellular Ca2+ fluxes. Train stimulus-induced Ca2+ influx (bottom) is simulated in the model by transiently increasing the Ca2+ influx rate constant (kinflux). Directly after reducing kinflux there is a rapid recovery leading to a prolonged Ca2+ plateau that readily fits the raw data (root mean square error: 2.3 × 10−8). Parameters for this fit are (± 95% confidence interval): Vmax, efflux: 6.2 ± 0.26 nm/s, kmax, uptake: 4.45 ± 0.82 (s−1), Vmax, release: 6.2 ± 1.14 nm/s. Unless otherwise stated, all subsequent model presentations were derived using the average values presented in Table 1. The values in Table 1 reflect parameters measured from changes to free Ca2+, before converting to rates of total Ca2+. B, For this and subsequent model graphs, Ca2+ levels reflect total, rather than free intracellular Ca2+ to correct for the presence of exogenous (fura and/or EGTA) and endogenous Ca2+ buffers present when estimating rates (see Materials and Methods). Serially reducing the rate constant of mitochondrial Ca2+ uptake (kmax, uptake) from 10 to 3.3 then 0 (s−1) attenuates the degree of mitochondrial Ca2+ uptake (right, light traces), slows the post-stimulus removal, and reduces CICR (left, light traces). C, Left inset, to include EGTA (0.5 mm), the bag cell neuron model of Ca2+ dynamics is increased to four components by adding Ca2+ removal by an exogenous Ca2+ binding agent. The clearance of Ca2+ by EGTA is determined by its forward (kon) and reverse (koff) rate constants, respectively. In the absence of EGTA, evoking Ca2+ influx causes a rise in Ca2+ and a subsequent Ca2+ plateau (left, dark trace). Under these conditions, mitochondrial Ca2+ increases then falls as Ca2+ is released into the cytosol (right, dark trace). The cytosolic Ca2+ response in the presence of EGTA slightly reduces peak Ca2+ influx magnitude and eliminates mitochondrial CICR (left, light traces). EGTA also partially attenuates the magnitude of mitochondrial Ca2+ influx after stimulation (right, light traces). D, In the presence of an EGTA component (0.5 mm), increasing the mitochondrial uptake rate constant (kmax, uptake) from 10 to 40 then 160 (s−1), potentiates the degree of mitochondrial Ca2+ loading (right, light traces), speeds the rate of cytosolic Ca2+ recovery after influx, and produces very limited CICR (left, light traces).

Transiently increasing the plasma membrane influx rate constant (kinflux), to replicate voltage-gated Ca2+ influx, produced a fast increase in cytosolic Ca2+, followed by a rapid recovery and a prolonged cytosolic Ca2+ plateau with a similar time course as that seen in actual bag cell neurons (Fig. 7B, left). Over the same time period, mitochondrial Ca2+ rapidly increased and decayed to prestimulus levels (Fig. 7B, right), corresponding to the uptake and release phases of cytosolic Ca2+, respectively. To demonstrate the necessity of mitochondrial Ca2+ uptake and release, Juptake was tempered by serially reducing the uptake rate constant (kmax,uptake). This produced a progressively smaller Ca2+ increase in the mitochondria, slowed the rate of cytosolic Ca2+ removal from peak influx, and reduced the CICR magnitude (Fig. 7B, left). These results are qualitatively similar to those seen in ours and other experiments when mitochondrial uptake is eliminated (Thayer and Miller, 1990; Friel and Tsien, 1994; Colegrove et al., 2000a; Geiger and Magoski, 2008).

Having established parameters that replicate our experiments, we sought to address the sensitivity of CICR to intracellular EGTA. The effects of EGTA on CICR could be due to competition with mitochondrial uptake for Ca2+ influx, causing a reduction in mitochondrial loading and subsequent release, or by EGTA binding the Ca2+ as it is extruded. To discern between these possibilities, a fourth component was added to the model, representing the buffering of Ca2+ by EGTA, which was determined by its forward (kon) and reverse reaction rate constants (koff; see Materials and Methods) (Nowycky and Pinter, 1993; Naraghi, 1997). Inclusion of an EGTA component (0.5 mm) in the compartment model caused a small reduction in the peak cytosolic Ca2+ rise during influx, while eliminating CICR (Fig. 7C, left). This in silico result is strikingly comparable to that observed in vitro (Fig. 1). In the EGTA-containing conditions, peak mitochondrial Ca2+ levels were slightly reduced following cytosolic Ca2+ influx, and consequently, decayed to prestimulus Ca2+ levels at a faster time course than in the absence of EGTA (Fig. 7C, right).

To determine whether increasing the degree of mitochondrial Ca2+ loading could rescue CICR in the presence of the EGTA component (0.5 mm), mitochondrial uptake was serially enhanced. Stepping the kmax, uptake from our standard value of 10 to 40 then 160 (s−1) increased mitochondrial Ca2+ to concentrations as large or higher than those seen in the absence of EGTA (compare Fig. 7C, right, D, right). Despite increasing the degree of mitochondrial Ca2+ available for release by enhancing mitochondrial uptake, CICR was only weakly rescued in the presence of EGTA. Thus, the sensitivity of CICR to EGTA is largely attributable to competition for Ca2+ released from the mitochondria.

Store-operated Ca2+ influx refills the ER Ca2+ store and is primarily cleared by the SERCA

In addition to voltage-gated Ca2+ influx and CICR, a third Ca2+ source prominent in the bag cell neurons is store-operated Ca2+ influx (Kachoei et al., 2006). It is well established that signaling cascades release ER Ca2+ through ryanodine- and inositol triphospate (IP3)-receptors in the bag cell neurons (Fink et al., 1988; Fisher et al., 1994; Geiger and Magoski, 2008). Here we evoked store-operated influx to determine whether a third Ca2+ source uses distinct removal mechanisms.

To measure store-operated Ca2+ influx, bag cell neurons were fura-filled by sharp-electrode pressure injection rather than recorded under whole-cell voltage clamp. This was possible because ER depletion and store-operated Ca2+ influx do not significantly alter the membrane potential, and therefore do not require voltage clamp (Kachoei et al., 2006). Store-operated influx was activated by applying 20 μm CPA in Ca2+-free medium to deplete the ER by inhibiting the SERCA (Seidler et al., 1989). After depletion and washout of CPA, the addition of Ca2+ back to the bath caused a rapid rise in intracellular Ca2+ as a result of Ca2+ influx through depletion-activated store-operated channels (Fig. 8A). A second application of CPA after influx and recovery elicited another rise in intracellular Ca2+ (Fig. 8A); however, its magnitude was significantly smaller than the first (Fig. 8B). In contrast, when store-operated influx recovery occurred in the presence of CPA (see below), a second application of CPA had no apparent effect on resting Ca2+ (data not shown).

Figure 8.

Figure 8.

The store-operated Ca2+ influx pathway is cleared by the SERCA to replete the ER. A, Addition of 20 μm CPA depletes ER Ca2+ in a cultured bag cell neuron pressure-injected with fura-PE3. Because neurons are not recorded under voltage clamp, a normal Na+-containing and K+-containing external solution is used. After the first depletion, CPA is washed out using bath exchange (at break); upon recording resumption, the addition of extracellular Ca2+ results in a rapid elevation of intracellular Ca2+. Delivering CPA a second time again evokes a Ca2+ rise, albeit smaller than the first, indicating repletion of the CPA-sensitive store. B, The magnitude of Ca2+ release to the second CPA exposure, after store-operated influx, is significantly smaller than the response elicited during the first CPA-induced depletion (unpaired Student's t test). C, Left, Washout of CPA, before addition of extracellular Ca2+, speeds the recovery of influx back to baseline. The removal of store-operated Ca2+ recovers at a slower rate in the presence of CPA (dark trace). Both neurons previously depleted in Ca2+-free ASW with 20 μm CPA. C, Right, Ca2+ influx, similar in size to that evoked by store-operated influx, caused by a short 5 Hz train of depolarizing stimuli from −80 to 0 mV in the presence (dark trace) and absence (light trace) of CPA. CPA has no effect on the speed of recovery from the short stimulus. D, Left, Mean percentage change in 340/380 (left) indicates that the increase in cytosolic Ca2+ during store-operated Ca2+ influx and the short train stimulus Ca2+ influx are not significantly different within and between conditions (ANOVA). D, Middle, CPA significantly increases the mean store-operated Ca2+ influx decay time constant (Welch corrected unpaired Student's t test) although having no effect on the mean τ of the similarly sized train stimulus-induced Ca2+ influx (unpaired Student's t test). D, Right, The percentage recovery at 5 min (right) after the application of extracellular Ca2+ is significantly shorter with prior CPA washout (Welch corrected unpaired Student's t test). However, the percentage recovery at 5 min after the short train stimulus Ca2+ influx is not significantly altered by CPA.

These results indicate that store-operated Ca2+ influx refills the ER via the SERCA. To determine the degree of Ca2+ removal by the SERCA, the recovery from peak store-operated Ca2+ influx was measured in the presence and absence of CPA. After ER depletion, washout of CPA and addition of extracellular Ca2+ evoked a transient rise in Ca2+ followed by a monoexponential decay, typically reaching pre-influx baseline within 10 min (Fig. 8C, left). If CPA remained in the bath after depletion (n = 17), store-operated influx was similar in size compared with controls (n = 17); however, Ca2+ recovery slowed dramatically, often requiring well over 15 min for full recovery (Fig. 8C, left). Furthermore, in CPA, following a brief initial recovery, Ca2+ levels often decayed to a new baseline at a Ca2+ concentration higher than pre-influx levels.

Our previous data established that the ER was ineffective at removing rapid voltage-gated Ca2+ influx. As the activity of the SERCA pump was demonstrated to be a relatively slow during store-operated influx, the effect of CPA on the recovery of a large Ca2+ influx may have gone undetected. Therefore, we produced a voltage-gated Ca2+ influx, sized-matched to the average store-operated influx, and tested its sensitivity to CPA (control n = 10, CPA n = 10). A brief (1.5–3 s) 5 Hz train stimulus from −80 to 0 mV applied to fura-loaded bag cell neurons under whole-cell voltage clamp elicited response magnitudes that were not significantly different than those during store-operated influx (Fig. 8D, left). Despite having the same absolute magnitude as store-operated influx, voltage-gated Ca2+ influx had a dramatically faster exponential recovery (Fig. 8D, middle). Consistent with this, the handling of voltage-gated Ca2+ influx showed no sensitivity to CPA, and rapidly declined to prestimulus levels within 2 min (Fig. 8C, right, D, middle, right).

Role for the mitochondria, but not the PMCA or Na+/Ca2+ exchanger, in store-operated Ca2+ influx

In other systems, mitochondria have been shown to be functionally coupled to store-operated influx (Gilabert and Parekh, 2000; Glitsch et al., 2002; Parekh and Putney, 2005; Naghdi et al., 2010). Therefore, influx was measured after 20 min pretreatment with 20 μm FCCP, to collapse the mitochondrial membrane potential. Compared with controls (n = 8), FCCP-treated neurons (n = 9) had a significantly reduced peak percentage change in response to bath application of Ca2+ after CPA-induced depletion (Fig. 9A, left, B). Furthermore, exposure to FCCP reduced the percentage of cells presenting a measurable influx signal. In the DMSO-treated group, 8 of 9 cells showed a response, whereas after FCCP, influx was observed only in 9 of 24 neurons (p < 0.02, Fisher's exact test). The contribution of acidic stores to the removal of store-operated influx was tested by exposing cells with 100 nm bafilomycin A (Fig. 9A, right). Incubation in bafilomycin A (n = 20) before the addition of extracellular Ca2+ did not change the peak store-operated influx (Fig. 9B), Ca2+ decay time constant (Fig. 9C), or the percentage recovery at 5 min post-peak Ca2+ compared with control (n = 16) (Fig. 9D).

Figure 9.

Figure 9.

Store-operated Ca2+ influx is reduced in magnitude by FCCP but is not influenced by acidic stores, the PMCA, or the Na+/Ca2+ exchanger. A, Left, Neurons exposed to 20 μm FCCP before the addition of extracellular Ca2+ (dark trace) causes a reduction in store-operated influx relative to control (light trace). A, Right, Exposure to 100 nm bafilomycin A (baf) before the addition of extracellular Ca2+ (dark trace) does not influence the size or recovery of store-operated Ca2+ influx. Neurons were previously depleted in Ca2+-free ASW with 20 μm CPA. B, Treatment with FCCP significantly reduces the peak Ca2+ influx after the addition of extracellular Ca2+ (left), whereas bafilomycin has no effect (both comparisons using Mann–Whitney U test). C, Summary data demonstrating that the mean decay time constant (τ) is significantly larger in CPA (Welch corrected unpaired Student's t test), but not in bafilomycin (unpaired Student's t test), 20 μm carboxyeosin (CE) (Welch corrected Student's t test), or where extracellular Na+ is replaced with NMDG (Mann–Whitney U test). D, CPA pretreatment significantly reduces the percentage recovery from peak Ca2+, whereas incubation with bafilomycin, carboxyeosin, or extracellular NMDG does not (all comparisons using unpaired Student's t test).

Contributions from the Na+/Ca2+ exchanger and the PMCA were examined by including extracellular NMDG in lieu of Na+ and 20 μm carboxyeosin, respectively. We used NMDG to inhibit the Na+/Ca2+ exchanger (Blaustein and Lederer, 1999; Zhang et al., 2004), rather than TEA, because TEA blocks K+ currents and would depolarize bag cell neurons in non-voltage-clamped conditions (Hagiwara and Saito, 1959; Kaczmarek and Strumwasser, 1984). Carboxyeosin, a different PMCA inhibitor (Shmigol et al., 1998), was used because the onset time of store-operated influx varied among individual cells during a recording, making proper timing of La3+ application impossible. Again, SERCA function was inhibited by 20 μm CPA following prior ER depletion. Blocking the Na+/Ca2+ exchanger (Na+ n = 12; NMDG n = 7) or the PMCA (control n = 12; carboxyeosin n = 16) did not significantly alter the decay time constants (Fig. 9C) or the percentage recovery at 5 min (Fig. 9D) despite the ongoing presence of CPA.

Discussion

Although neuron-specific or compartment-specific expression of Ca2+ clearance systems has been reported, there is sparse evidence of differential handling mechanisms between Ca2+ sources. White and Reynolds (1995) found that a glutamate-evoked Ca2+ response is more sensitive to the disruption of the Na+/Ca2+ exchanger and mitochondria than voltage-gated Ca2+ influx. Conversely, in mouse taste receptor cells voltage-gated Ca2+ influx, but not Ca2+ release, is Na+/Ca2+ exchanger sensitive (Szebenyi et al., 2010). This work is the first to fully characterize the disparate contribution of handling mechanisms to multiple Ca2+ sources in a single neuronal species. We demonstrate: (1) voltage-gated Ca2+ influx is primarily removed by the mitochondria with a secondary contribution from the PMCA; (2) CICR arises from subsequent mitochondrial Ca2+ release, which is then handled by the Na+/Ca2+ exchanger and the PMCA; and (3) store-operated Ca2+ influx is sequestered via the SERCA (Fig. 10). This is profound, given that each Ca2+ source provides unique contributions to neuronal function.

Figure 10.

Figure 10.

Ca2+ from multiple sources is cleared by distinct sets of Ca2+ removal systems in the bag cell neurons. A conceptual model of Ca2+ dynamics during an afterdischarge based on the current study and prior work by our laboratory and others (Fink et al., 1988; Fisher et al., 1994; Michel and Wayne, 2002; Kachoei et al., 2006; Geiger and Magoski, 2008). Top, right inset, A sample trace of intracellular Ca2+ presumed to reflect afterdischarge dynamics in vivo. Numbers correspond to a series of chronological events depicted in the main illustration: 1, During the fast phase of the afterdischarge there is a large Ca2+ influx through action potential-evoked voltage-gated Ca2+ channels (VGCC). 2, This early Ca2+ influx is predominantly removed by rapid mitochondrial uptake, with ancillary assistance from the PMCA. 3, Ca2+ accumulates in the mitochondria, after which it is slowly extruded into the cytosol through a TPP-sensitive Ca2+-exchanger. Additionally, second messenger cascades are activated that initiate Ca2+ liberation from the ER through IP3 and ryanodine receptors. 4, The release of Ca2+ from intracellular stores causes a prolonged rise in cytosolic Ca2+ that is removed by the Na+/Ca2+ exchanger and the PMCA. 5, Last, sustained Ca2+ release during the afterdischarge depletes the ER and initiates store-operated Ca2+ influx: a distinct Ca2+ source that is removed by the SERCA into the ER lumen.

Voltage-gated Ca2+ entry is the primary means to increase cytosolic Ca2+ (Catterall and Few, 2008). In the bag cell neurons, this gates Ca2+-dependent cation channels that promote the afterdischarge (Whim and Kaczmarek, 1998; Lupinsky and Magoski, 2006; Hung and Magoski, 2007). We show that voltage-gated Ca2+ influx is handled by the mitochondria, similar to some neurons and neuroendocrine cells (Werth and Thayer, 1994; Kaftan et al., 2000). However, unlike chromaffin cells, where mitochondrial uptake is engaged only at high Ca2+ (Herrington et al., 1996), bag cell neuron mitochondria function near resting and peak Ca2+ levels. Neuroendocrine cells provide hormones to orchestrate fundamental behaviors, such as feeding, drinking, and reproduction (Arnauld et al., 1974; Lincoln and Wakerley, 1974; Kawakami et al., 1982). Considering the importance of energy status to these behaviors, it is unsurprising that mitochondria are the first recipients of Ca2+ during bag cell neuron excitation. Essentially, mitochondria are situated to act as gatekeepers of the afterdischarge and reproductive status.

In the bag cell neurons, a 1 min stimulus evoked CICR that is sensitive to TPP or intracellular EGTA. A compartment model of bag cell neuron Ca2+ suggests that the effect of EGTA on CICR was partially due to a reduction in mitochondrial Ca2+ loading, but mainly the result of EGTA binding extruded mitochondrial Ca2+. Likely, this is possible because of slow Ca2+ release versus fast uptake across the mitochondria. Similarly, maximal uptake in sympathetic neuron mitochondria approaches ∼120 nm/s, whereas release is far slower (∼35 nm/s) (Colegrove et al., 2000a). Such kinetics render mitochondrial Ca2+ release sensitive to competitive cytoplasmic buffers. CICR transduces short-lived events into protracted Ca2+ signals, which impact biochemical pathways more effectively than the initial response alone. For example, CICR promotes posttetanic transmitter release at mouse neuromuscular junctions and hippocampal synapses (Garcia-Chacon et al., 2006; Lee et al., 2007), and prolongs Ca2+-dependent transcription in dorsal root ganglion neurons (Kim and Usachev, 2009). During an afterdischarge, CICR may serve analogously to sustain hormone release. Indeed, some ELH secretion is independent of extracellular Ca2+, implicating the involvement of Ca2+ stores (Wayne et al., 1998; Michel and Wayne, 2002). The prolonged duration of bag cell neuron CICR reflects a combination of persistent Ca2+ release, and removal by the PMCA and Na+/Ca2+ exchanger, providing an efficient form of recycling Ca2+ to the extracellular space. These extrusion mechanisms, which are slow (∼4 nm/s), may shape CICR but not necessarily dampen its effectiveness. The mitochondria cope with Ca2+ influx and transduce it into a lengthy release that is extruded without saturating the maximal rate of plasma membrane transport.

Store-operated Ca2+ influx was exclusively handled by the SERCA. Accordingly, this pump refilled the ER, consistent with similar roles of store-operated Ca2+ influx in other cells (Hoth and Penner, 1992; Putney, 2003). In the bag cell neurons, ER Ca2+ release occurs via IP3 and ryanodine receptors, and can activate a BK-like K+ channel and cation current (Fink et al., 1988; Knox et al., 1996) and may contribute to ELH secretion (Michel and Wayne, 2002). Activation of store-operated influx requires ER depletion. As such, it likely presents a delayed activation, once the majority of Ca2+ liberation has occurred (Fig. 10), and could sustain processes activated by ER Ca2+. Store-operated influx was also reduced when mitochondrial function was absent. This is also seen in T-lymphocytes and retinoblastoma-1 cells, where mitochondrial Ca2+ sequestration prevents Ca2+-dependent inhibition of store-operated channels (Glitsch et al., 2002; Naghdi et al., 2010). In the bag cell neurons, although neither the Na+/Ca2+ exchanger nor the PMCA were involved, there was a residual recovery from peak store-operated influx when the SERCA was inhibited. Therefore, a similar explanation involving mitochondrial Ca2+ clearance is plausible.

The specificity between Ca2+ sources and clearance mechanisms seen in this study could be attributed to the duration of Ca2+ exposure, the affinity of the Ca2+ handling systems, and/or the distribution of the buffers and influx sources (Gabso et al., 1997; Berridge et al., 2003). The proportion of removal by each handling mechanism depends on influx magnitude. For example, the PMCA and SERCA are reported to have high Ca2+-binding affinities and remove small Ca2+ loads, whereas the low affinity Na+/Ca2+ exchanger and mitochondria are engaged only by high intracellular Ca2+ (Herrington et al., 1996; Blaustein and Lederer, 1999; Berridge et al., 2003). Such is not the case here, where voltage-gated Ca2+ influx of similar size as store-operated influx or CICR does not engage the Na+/Ca2+ exchanger or SERCA. This is also confirmed by our prior work demonstrating the CPA-insensitivity of a small voltage-gated Ca2+ influx elicited in sharp-electrode recorded bag cell neurons (Geiger and Magoski, 2008). It is doubtful that differences in Ca2+ signal kinetics contributed to our results. A prolonged poststimulus Ca2+ influx showed no sensitivity to the Na+/Ca2+ exchanger, despite demonstrating a similar duration as CICR. Furthermore, CICR occurred with a duration and magnitude comparable to store-operated Ca2+ influx, and yet presented a different involvement of the SERCA, PMCA, and Na+/Ca2+ exchanger. Thus, the differential control we observe seems attributable to variation in the location of Ca2+ sources and/or handling systems. Indeed, Ca2+ channel types, including those in the bag cell neurons, can occupy discrete areas within the soma (White and Kaczmarek, 1997). Furthermore, in cortical neurons and astrocytes, the Na+/Ca2+ exchanger parallels the position of intracellular organelles and can occupy distinct regions relative to the PMCA (Juhaszova et al., 1996, 2000). Organelles can also occupy unique cellular loci. In bullfrog sympathetic neurons, SERCA clusters near the plasma membrane, whereas the mitochondria form a centralized inner ring (McDonough et al., 2000). Preferential localization can have functional implications, such as in cervical ganglion neurons, where Ca2+ handling is biased toward a given source, by the ER and mitochondria associating with Cav2 but not Cav1 channels (Wheeler et al., 2012).

Unlike in the calyx of Held and retinal bipolar cells (Zenisek and Mathews, 2000; Kim et al., 2005), we find that, in the absence of a dominant removal system, the contribution of an otherwise uninvolved clearance mechanism is not enhanced. Compensation likely occurs because Ca2+ gradients disperse after influx (Hua et al., 1993), diffusing to buffers that are displaced from the source. Therefore, even when considering a heterogeneous distribution of removal systems, our results are surprising. Potentially, Ca2+ signals are compartmentalized by diffusional barriers such as partitions formed by membrane junctions (e.g., ER-plasma membrane) (Carrasco and Meyer, 2011) or “Ca2+-sponging” by immobile Ca2+-binding proteins (Nowycky and Pinter, 1993; Weiss et al., 2012), which are found in other Aplysia neurons. Estimates of the Ca2+-diffusion coefficient in cultured Aplysia neurons (≤16 μm2/s) are similar to those of specialized Ca2+ signaling structures, such as dendrites (10–50 μm2/s) or photoreceptor outer segments (15 μm2/s) (Gabso et al., 1997; Murthy et al., 2000; Nakatani et al., 2002). Therefore, auxiliary systems may have contributed to removal, but their effects were slow or delayed.

Although differential clearance between sources may reflect an efficient form of Ca2+ homeostasis or Ca2+-metabolism coupling (Rizzuto et al., 1998; Chouhan et al., 2012), a more intriguing possibility is that it shapes Ca2+ signals to determine which Ca2+-dependent pathway is activated. In the aforementioned superior cervical ganglion neurons, the mitochondria and ER preferentially remove Ca2+ influx from Cav2, but not Cav1 channels, permitting Cav1-specific activation of transcription (Wheeler et al., 2012). For bag cell neurons, Ca2+ entry through a cation channel, but not voltage-gated Ca2+ influx or Ca2+ release, induces a period of refractoriness to further stimulation (Magoski et al., 2000). Furthermore, following the afterdischarge, there is a Ca2+-influx-dependent increase in ELH synthesis, which could be brought about by differences in the clearance of Ca2+ from distinct sources (Wayne et al., 2004). Differential handling between Ca2+ sources could underlie the ability of discrete Ca2+ sources to activate specific downstream targets. Our work reinforces the shifting view of Ca2+ buffers as passive determinants of intracellular Ca2+ to dynamic regulators that can control excitability, secretion, synaptic plasticity, and gene expression (Krizaj and Copenhagen, 1998; Holthoff et al., 2002; Kim and Usachev, 2009; Simons et al., 2009).

Footnotes

This work was supported by Canadian Institutes of Health Research (CIHR) and Natural Sciences and Engineering Research Council of Canada (NSERC) operating grants to N.S.M. We thank H. M. Hodgson and S. L. Smith for technical assistance. C.J.G. holds a NSERC Post-Graduate Scholarship, and N.S.M. holds a CIHR New Investigator Award.

References

  1. Arch S. Polypeptide secretion from the isolated parietovisceral ganglion of Aplysia californica. J Gen Physiol. 1972;59:47–59. doi: 10.1085/jgp.59.1.47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Armstrong CM, Hille B. Voltage-gated ion channels and electrical excitability. Neuron. 1998;20:371–380. doi: 10.1016/S0896-6273(00)80981-2. [DOI] [PubMed] [Google Scholar]
  3. Arnauld E, Vincent JD, Dreifuss JJ. Firing patterns of hypothalamic supraoptic neurons during water deprivation in monkeys. Science. 1974;185:535–537. doi: 10.1126/science.185.4150.535. [DOI] [PubMed] [Google Scholar]
  4. Babcock DF, Herrington J, Goodwin PC, Park YB, Hille B. Mitochondrial participation in the intracellular Ca2+ network. J Cell Biol. 1997;136:833–844. doi: 10.1083/jcb.136.4.833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signaling. Nat Rev Mol Cell Biol. 2000;1:11–21. doi: 10.1038/35036035. [DOI] [PubMed] [Google Scholar]
  6. Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4:517–529. doi: 10.1038/nrm1155. [DOI] [PubMed] [Google Scholar]
  7. Blaustein MP, Lederer JW. Sodium/calcium exchange: its physiological implications. Physiol Rev. 1999;79:763–854. doi: 10.1152/physrev.1999.79.3.763. [DOI] [PubMed] [Google Scholar]
  8. Bowman EJ, Siebers A, Altendorf K. Bafilomycins: a class of inhibitors of membrane ATPases from microorganisms, animal cells, and plant cells. Proc Natl Acad Sci U S A. 1988;85:7972–7976. doi: 10.1073/pnas.85.21.7972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Carafoli E. Calcium pump of the plasma membrane. Physiol Rev. 1991;71:129–153. doi: 10.1152/physrev.1991.71.1.129. [DOI] [PubMed] [Google Scholar]
  10. Caride AJ, Filoteo AG, Penheiter AR, Pászty K, Enyedi A, Penniston JT. Delayed activation of the plasma membrane calcium pump by a sudden increase in Ca2+: fast pumps reside in fast cells. Cell Calcium. 2001;30:49–57. doi: 10.1054/ceca.2001.0212. [DOI] [PubMed] [Google Scholar]
  11. Carrasco S, Meyer T. STIM proteins and the endoplasmic reticulum-plasma membrane junctions. Annu Rev Biochem. 2011;80:973–1000. doi: 10.1146/annurev-biochem-061609-165311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Catterall WA, Few AP. Calcium channel regulation and presynaptic plasticity. Neuron. 2008;59:882–901. doi: 10.1016/j.neuron.2008.09.005. [DOI] [PubMed] [Google Scholar]
  13. Chouhan AK, Ivannikov MV, Lu Z, Sugimori M, Llinas RR, Macleod GT. Cytosolic calcium coordinates mitochondrial energy metabolism with presynaptic activity. J Neurosci. 2012;32:1233–1243. doi: 10.1523/JNEUROSCI.1301-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Christensen KA, Myers JT, Swanson JA. pH-dependent regulation of lysosomal calcium in macrophages. J Cell Sci. 2002;115:599–607. doi: 10.1242/jcs.115.3.599. [DOI] [PubMed] [Google Scholar]
  15. Clapham DE. Calcium signaling. Cell. 1995;80:259–268. doi: 10.1016/0092-8674(95)90408-5. [DOI] [PubMed] [Google Scholar]
  16. Colegrove SL, Albrecht MA, Friel DD. Dissection of mitochondrial Ca2+ uptake and release fluxes in situ after depolarization evoked [Ca2+]i elevations in sympathetic neurons. J Gen Physiol. 2000a;115:351–370. doi: 10.1085/jgp.115.3.351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Colegrove SL, Albrecht MA, Friel DD. Quantitative analysis of mitochondrial Ca2+uptake and release pathways in sympathetic neurons. Reconstruction of the recovery after depolarization-evoked [Ca2+]i elevations. J Gen Physiol. 2000b;115:381–388. doi: 10.1085/jgp.115.3.371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Deisseroth K, Heist EK, Tsien RW. Translocation of calmodulin to the nucleus supports CREB phosphorylation in hippocampal neurons. Nature. 1998;392:198–202. doi: 10.1038/32448. [DOI] [PubMed] [Google Scholar]
  19. DeRiemer SA, Kaczmarek LK, Lai Y, McGuinness TL, Greengard P. Calcium/calmodulin-dependent protein phosphorylation in the nervous system of Aplysia. J Neurosci. 1984;4:1618–1625. doi: 10.1523/JNEUROSCI.04-06-01618.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Fierro L, DiPolo R, Larno I. Intracellular calcium clearance in Purkinje cell somata from rat cerebellar slices. J Physiol. 1998;510:499–512. doi: 10.1111/j.1469-7793.1998.499bk.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Fink LA, Connor JA, Kaczmarek LK. Inositol triphosphate releases intracellularly stored calcium and modulates ion channels in molluscan neurons. J Neurosci. 1988;8:2544–2555. doi: 10.1523/JNEUROSCI.08-07-02544.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Fisher TE, Levy S, Kaczmarek LK. Transient changes in intracellular calcium associated with a prolonged increase in excitability in neurons of Aplysia californica. J Neurophysiol. 1994;71:1254–1257. doi: 10.1152/jn.1994.71.3.1254. [DOI] [PubMed] [Google Scholar]
  23. Friel DD, Tsien RW. An FCCP-sensitive Ca2+ store in bullfrog sympathetic neurons and its participation in stimulus-evoked changes in [Ca2+]i. J Neurosci. 1994;14:4007–4024. doi: 10.1523/JNEUROSCI.14-07-04007.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Gabso M, Neher E, Spira ME. Low mobility of the Ca2+ buffers in axons of cultured Aplysia neurons. Neuron. 1997;18:473–481. doi: 10.1016/S0896-6273(00)81247-7. [DOI] [PubMed] [Google Scholar]
  25. García-Chacón LE, Nguyen KT, David G, Barrett EF. Extrusion of Ca2+ from mouse motor terminal mitochondria via a Na+-Ca2+ exchanger increases post-tetanic evoked release. J Physiol. 2006;574:663–675. doi: 10.1113/jphysiol.2006.110841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Gardam KE, Geiger JE, Hickey CM, Hung AY, Magoski NS. Flufenamic acid affects multiple currents and causes intracellular Ca2+ release in Aplysia bag cell neurons. J Neurophysiol. 2008;100:38–49. doi: 10.1152/jn.90265.2008. [DOI] [PubMed] [Google Scholar]
  27. Geiger JE, Magoski NS. Ca2+-induced Ca2+-release in Aplysia bag cell neurons requires interaction between mitochondrial and endoplasmic reticulum stores. J Neurophysiol. 2008;100:24–37. doi: 10.1152/jn.90356.2008. [DOI] [PubMed] [Google Scholar]
  28. Gilabert JA, Parekh AB. Respiring mitochondria determine the pattern of activation and inactivation of the store-operated Ca2+ current ICRAC. EMBO J. 2000;19:6401–6407. doi: 10.1093/emboj/19.23.6401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Glitsch MD, Bakowski D, Parekh AB. Store-operated Ca2+ entry depends on mitochondrial Ca2+ uptake. EMBO J. 2002;21:6744–6754. doi: 10.1093/emboj/cdf675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Gonçalves PP, Meireles SM, Neves P, Vale MG. Synaptic vesicle Ca2+/H+ antiport: dependence on the proton electrochemical gradient. Mol Brain Res. 1999;71:178–184. doi: 10.1016/S0169-328X(99)00183-7. [DOI] [PubMed] [Google Scholar]
  31. Gorman AL, Thomas MV. Intracellular calcium accumulation during depolarization in a molluscan neuron. J Physiol. 1980;308:259–285. doi: 10.1113/jphysiol.1980.sp013471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Grynkiewicz G, Poenie M, Tsien RY. A new generation of Ca2+ indicators with greatly improved fluorescence properties. J Biol Chem. 1985;260:3440–3450. [PubMed] [Google Scholar]
  33. Gunter KK, Gunter TE. Transport of calcium by mitochondria. J Bioenerg Biomembr. 1994;26:471–485. doi: 10.1007/BF00762732. [DOI] [PubMed] [Google Scholar]
  34. Gunter TE, Pfeiffer DR. Mechanisms by which mitochondria transport calcium. Am J Physiol. 1990;258:755–786. doi: 10.1152/ajpcell.1990.258.5.C755. [DOI] [PubMed] [Google Scholar]
  35. Hagiwara S, Saito N. Voltage-current relations in nerve cell membrane of Onchidium verruculatum. J Physiol. 1959;148:161–179. doi: 10.1113/jphysiol.1959.sp006279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Herrington J, Park YB, Babcock DF, Hille B. Dominant role of mitochondria in clearance of large Ca2+ loads from rat adrenal chromaffin cells. Neuron. 1996;16:219–228. doi: 10.1016/S0896-6273(00)80038-0. [DOI] [PubMed] [Google Scholar]
  37. Heytler PG, Prichard WW. A new class of uncoupling agents–carbonyl cyanide phenylhydrazones. Biochem Biophys Res Commun. 1962;7:272–275. doi: 10.1016/0006-291X(62)90189-4. [DOI] [PubMed] [Google Scholar]
  38. Hickey CM, Geiger JE, Groten CJ, Magoski NS. Mitochondrial Ca2+ activates a cation current in Aplysia bag cell neurons. J Neurophysiol. 2010;103:1543–1556. doi: 10.1152/jn.01121.2009. [DOI] [PubMed] [Google Scholar]
  39. Holthoff K, Tsay D, Yuste R. Calcium dynamics of spines depend on their dendritic location. Neuron. 2002;33:425–437. doi: 10.1016/S0896-6273(02)00576-7. [DOI] [PubMed] [Google Scholar]
  40. Hoth M, Penner R. Depletion of intracellular calcium stores activates a calcium current in mast cells. Nature. 1992;355:353–356. doi: 10.1038/355353a0. [DOI] [PubMed] [Google Scholar]
  41. Hua SY, Nohmi M, Kuba K. Characteristics of Ca2+ release induced by Ca2+ influx in cultured bullfrog sympathetic neurons. J Physiol. 1993;464:245–272. doi: 10.1113/jphysiol.1993.sp019633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hung AY, Magoski NS. Activity-dependent initiation of a prolonged depolarization in Aplysia bag cell neurons: role for a cation channel. J Neurophysiol. 2007;97:2465–2479. doi: 10.1152/jn.00941.2006. [DOI] [PubMed] [Google Scholar]
  43. Jeon D, Yang YM, Jeong MJ, Philipson KD, Rhim Hyewhon, Shin HS. Enhanced learning and memory in mice lacking Na+/Ca2+ exchanger 2. Neuron. 2003;38:965–976. doi: 10.1016/S0896-6273(03)00334-9. [DOI] [PubMed] [Google Scholar]
  44. Jonas EA, Knox RJ, Smith TC, Wayne NL, Connor JA, Kaczmarek LK. Regulation by insulin of a unique neuronal Ca2+ pool and neuropeptide secretion. Nature. 1997;385:343–346. doi: 10.1038/385343a0. [DOI] [PubMed] [Google Scholar]
  45. Juhaszova M, Shimizu H, Borin ML, Yip RK, Santiago EM, Lindenmayer GE, Blaustein MP. Localization of the Na+/Ca2+ exchanger in vascular smooth muscle, and in neurons and astrocytes. Ann N Y Acad Sci. 1996;779:318–335. doi: 10.1111/j.1749-6632.1996.tb44804.x. [DOI] [PubMed] [Google Scholar]
  46. Juhaszova M, Church P, Blaustein MP, Stanley EF. Location of calcium transporters at presynaptic terminals. Eur J Neurosci. 2000;12:839–846. doi: 10.1046/j.1460-9568.2000.00974.x. [DOI] [PubMed] [Google Scholar]
  47. Kachoei BA, Knox RJ, Uthuza D, Levy S, Kaczmarek LK, Magoski NS. A store-operated Ca2+ influx pathway in the bag cell neurons of Aplysia. J Neurophysiol. 2006;96:2688–2698. doi: 10.1152/jn.00118.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Kaczmarek LK, Strumwasser F. A voltage-clamp analysis of currents in isolated peptidergic neurons of Aplysia. J Neurophysiol. 1984;52:340–349. doi: 10.1152/jn.1984.52.2.340. [DOI] [PubMed] [Google Scholar]
  49. Kaczmarek LK, Jennings KR, Strumwasser F. An early sodium and a late calcium phase in the afterdischarge of peptide-secreting neurons of Aplysia. Brain Res. 1982;238:105–115. doi: 10.1016/0006-8993(82)90774-0. [DOI] [PubMed] [Google Scholar]
  50. Kaftan EJ, Xu T, Abercrombie RF, Hille B. Mitochondria shape hormonally induced cytoplasmic calcium oscillations and modulate exocytosis. J Biol Chem. 2000;18:25465–25470. doi: 10.1074/jbc.M000903200. [DOI] [PubMed] [Google Scholar]
  51. Karadjov JS, Kudzina LYu, Zinchenko VP. TPP+ inhibits Na+ stimulated Ca2+ efflux from brain mitochondria. Cell Calcium. 1986;7:115–119. doi: 10.1016/0143-4160(86)90014-X. [DOI] [PubMed] [Google Scholar]
  52. Kawakami M, Uemura T, Hayashi R. Electrophysiological correlates of pulsatile gonadotropin release in rats. Neuroendocrinology. 1982;55:63–67. doi: 10.1159/000123356. [DOI] [PubMed] [Google Scholar]
  53. Kim MH, Lee SH, Park KH, Ho WK, Lee SH. Distribution of K+-dependent Na+/Ca2+ exchangers in the rat supraoptic magnocellular neuron is polarized to axon terminals. J Neurosci. 2003;23:11673–11680. doi: 10.1523/JNEUROSCI.23-37-11673.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Kim MH, Korogod N, Schneggenburger R, Ho WK, Lee SH. Interplay between Na+/Ca2+ exchangers and mitochondria in Ca2+ clearance at the calyx of held. J Neurosci. 2005;25:6057–6065. doi: 10.1523/JNEUROSCI.0454-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Kim MS, Usachev YM. Mitochondrial Ca2+ cycling facilitates activation of the transcription factor NFAT in sensory neurons. J Neurosci. 2009;29:12101–12114. doi: 10.1523/JNEUROSCI.3384-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Knox RJ, Jonas EA, Kao LS, Smith PJ, Connor JA, Kaczmarek LK. Ca2+ influx and activation of a cation current are coupled to intracellular Ca2+ release in peptidergic neurons of Aplysia californica. J Physiol. 1996;494:627–639. doi: 10.1113/jphysiol.1996.sp021520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Krizaj D, Copenhagen DR. Compartmentalization of calcium extrusion mechanisms in the outer and inner segments of photoreceptors. Neuron. 1998;21:249–256. doi: 10.1016/S0896-6273(00)80531-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Kupfermann I, Kandel ER. Electrophysiological properties and functional interconnections of two symmetrical neurosecretory clusters (bag cells) in abdominal ganglion of Aplysia. J Neurophysiol. 1970;33:865–876. doi: 10.1152/jn.1970.33.6.865. [DOI] [PubMed] [Google Scholar]
  59. Lee D, Lee KH, Ho WK, Lee SH. Target cell-specific involvement of presynaptic mitochondria in post-tetanic potentiation at hippocampal mossy fiber synapses. J Neurosci. 2007;27:13603–13613. doi: 10.1523/JNEUROSCI.3985-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Lincoln DW, Wakerley JB. Electrophysiological evidence for the activation of supraoptic neurones during the release of oxytocin. J Physiol. 1974;242:533–554. doi: 10.1113/jphysiol.1974.sp010722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Loechner KJ, Azhderian EM, Dreyer R, Kaczmarek LK. Progressive potentiation of peptide release during a neuronal discharge. J Neurophysiol. 1990;63:738–744. doi: 10.1152/jn.1990.63.4.738. [DOI] [PubMed] [Google Scholar]
  62. Lupinsky DA, Magoski NS. Ca2+-dependent regulation of a non-selective cation channel from Aplysia bag cell neurones. J Physiol. 2006;575:491–506. doi: 10.1113/jphysiol.2006.105833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Magoski NS, Knox RJ, Kaczmarek LK. Activation of a Ca2+-permeable cation channel produces a prolonged attenuation of intracellular Ca2+ release in Aplysia bag cell neurones. J Physiol. 2000;522:271–283. doi: 10.1111/j.1469-7793.2000.t01-2-00271.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. McDonough SI, Cseresnyes Z, Schneider MF. Origin sites of calcium release and calcium oscillations in frog sympathetic neurons. J Neurosci. 2000;20:9059–9070. doi: 10.1523/JNEUROSCI.20-24-09059.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Michel S, Wayne NL. Neurohormone secretion persists after post-afterdischarge membrane depolarization and cytosolic calcium elevation in peptidergic neurons in intact nervous tissue. J Neurosci. 2002;22:9063–9069. doi: 10.1523/JNEUROSCI.22-20-09063.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Morgans CW, El Far O, Berntson A, Wässle H, Taylor WR. Calcium extrusion from mammalian photoreceptor terminals. J Neurosci. 1998;18:2467–2474. doi: 10.1523/JNEUROSCI.18-07-02467.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Murthy VN, Sejnowski TJ, Stevens CF. Dynamics of dendritic calcium transients evoked by quantal release at excitatory hippocampal synapses. Proc Natl Acad Sci U S A. 2000;97:901–906. doi: 10.1073/pnas.97.2.901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Naghdi S, Waldeck-Weiermair M, Fertschai I, Poteser M, Graier WF, Malli R. Mitochondrial Ca2+ uptake and not mitochondrial motility is required for STIM1-Orai1-dependent store-operated Ca2+ entry. J Cell Sci. 2010;123:2553–2564. doi: 10.1242/jcs.070151. [DOI] [PubMed] [Google Scholar]
  69. Nakatani K, Chen C, Koutalos Y. Calcium diffusion coefficient in rod photoreceptor outer segments. Biophys J. 2002;82:728–739. doi: 10.1016/S0006-3495(02)75435-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Naraghi M. T-jump study of calcium binding kinetics of calcium chelators. Cell Calcium. 1997;22:255–268. doi: 10.1016/S0143-4160(97)90064-6. [DOI] [PubMed] [Google Scholar]
  71. Neering IR, McBurney RN. Role for microsomal Ca2+ storage in mammalian neurones? Nature. 1984;309:158–160. doi: 10.1038/309158a0. [DOI] [PubMed] [Google Scholar]
  72. Neher E, Augustine GJ. Calcium gradients and buffers in bovine chromaffin cells. J Physiol. 1992;450:273–301. doi: 10.1113/jphysiol.1992.sp019127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Nowycky MC, Pinter MJ. Time courses of calcium and calcium-bound buffers following calcium influx in a model cell. Biophys J. 1993;64:77–91. doi: 10.1016/S0006-3495(93)81342-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Parekh AB, Putney JW., Jr Store-operated calcium channels. Physiol Rev. 2005;85:757–810. doi: 10.1152/physrev.00057.2003. [DOI] [PubMed] [Google Scholar]
  75. Pinsker HM, Dudek FE. Bag cell control of egg-laying in freely behaving Aplysia. Science. 1977;197:490–493. doi: 10.1126/science.197.4302.490. [DOI] [PubMed] [Google Scholar]
  76. Putney JW., Jr Capacitative calcium entry in the nervous system. Cell Calcium. 2003;34:339–344. doi: 10.1016/S0143-4160(03)00143-X. [DOI] [PubMed] [Google Scholar]
  77. Rizzuto R, Pozzan T. Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol Rev. 2006;86:369–408. doi: 10.1152/physrev.00004.2005. [DOI] [PubMed] [Google Scholar]
  78. Rizzuto R, Pinton P, Carrington W, Fay FS, Fogarty KE, Lifshitz LM, Tuft RA, Pozzan T. Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses. Science. 1998;280:1763–1766. doi: 10.1126/science.280.5370.1763. [DOI] [PubMed] [Google Scholar]
  79. Sanchez-Armass S, Blaustein MP. Role of sodium-calcium exchange in regulation of intracellular calcium in nerve terminals. Am J Physiol. 1987;252:595–603. doi: 10.1152/ajpcell.1987.252.6.C595. [DOI] [PubMed] [Google Scholar]
  80. Seidler NW, Jona I, Vegh M, Martonosi A. Cyclopiazonic acid is a specific inhibitor of the Ca2+-ATPase of sarcoplasmic reticulum. J Biol Chem. 1989;264:17816–17823. [PubMed] [Google Scholar]
  81. Shmigol A, Eisner DA, Wray S. Carboxyeosin decreases the rate of decay of the [Ca2+]i transient in uterine smooth muscle cells isolated from pregnant rats. Pflugers Arch. 1998;437:158–160. doi: 10.1007/s004240050761. [DOI] [PubMed] [Google Scholar]
  82. Simons SB, Escobedo Y, Yasuda R, Dudek SM. Regional differences in hippocampal calcium handling provide a cellular mechanism for limiting plasticity. Proc Natl Acad Sci U S A. 2009;106:14080–14084. doi: 10.1073/pnas.0904775106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Smith SJ, MacDermott AB, Weight FF. Detection of intracellular Ca2+ transients in sympathetic neurones using arsenazo III. Nature. 1983;304:350–352. doi: 10.1038/304350a0. [DOI] [PubMed] [Google Scholar]
  84. Szebenyi SA, Laskowski AI, Medler KF. Sodium/calcium exchangers selectively regulate calcium signaling in mouse taste receptor cells. J Neurophysiol. 2010;104:529–538. doi: 10.1152/jn.00118.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Tam AK, Geiger JE, Hung AY, Groten CJ, Magoski NS. Persistent Ca2+ current contributes to a prolonged depolarization in Aplysia bag-cell neurons. J Neurophysiol. 2009;102:3753–3765. doi: 10.1152/jn.00669.2009. [DOI] [PubMed] [Google Scholar]
  86. Tam AK, Gardam KE, Lamb S, Kachoei BA, Magoski NS. Role for protein kinase C in controlling Aplysia bag cell neuron excitability. Neuroscience. 2011;179:41–55. doi: 10.1016/j.neuroscience.2011.01.037. [DOI] [PubMed] [Google Scholar]
  87. Tang Y, Zucker RS. Mitochondrial involvement in post-tetanic potentiation of synaptic transmission. Neuron. 1997;18:483–491. doi: 10.1016/S0896-6273(00)81248-9. [DOI] [PubMed] [Google Scholar]
  88. Thayer SA, Miller RJ. Regulation of the intracellular free calcium concentration in single dorsal root ganglion neurones in vitro. J Physiol. 1990;425:85–115. doi: 10.1113/jphysiol.1990.sp018094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Tidow H, Poulsen LR, Andreeva A, Knudsen M, Hein KL, Wiuf C, Palmgren MG, Nissen P. A bimodular mechanism of calcium control in eukaryotes. Nature. 2012;491:468–472. doi: 10.1038/nature11539. [DOI] [PubMed] [Google Scholar]
  90. Vorndran C, Minta A, Poenie M. New fluorescent calcium indicators designed for cytosolic retention or measuring calcium near membranes. J Biophys. 1995;69:2112–2124. doi: 10.1016/S0006-3495(95)80082-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Wayne NL, Kim J, Lee E. Prolonged hormone secretion from neuroendocrine cells of Aplysia is independent of extracellular Ca2+ J Neuroendocrinol. 1998;10:529–537. doi: 10.1046/j.1365-2826.1998.00235.x. [DOI] [PubMed] [Google Scholar]
  92. Wayne NL, Lee W, Michel S, De Quintana SB. Post-afterdischarge depolarization does not stimulate prolonged neurohormone secretion but is required for activity-dependent stimulation of neurohormone biosynthesis from peptidergic neurons. Endocrinology. 2004;145:1678–1684. doi: 10.1210/en.2003-1447. [DOI] [PubMed] [Google Scholar]
  93. Weiss S, Kohn E, Dadon D, Katz B, Peters M, Lebendiker M, Kosloff M, Colley NJ, Minke B. Compartmentalization and Ca2+ buffering are essential for prevention of light-induced retinal degeneration. J Neurosci. 2012;32:14696–14708. doi: 10.1523/JNEUROSCI.2456-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Werth JL, Thayer SA. Mitochondria buffer physiological calcium loads in cultured rat dorsal root ganglion neurons. J Neurosci. 1994;14:348–356. doi: 10.1523/JNEUROSCI.14-01-00348.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Wheeler DG, Groth RD, Ma H, Barrett CF, Owen SF, Safa P, Tsien RW. Cav1 and Cav2 channels engage distinct modes of Ca2+ signaling to control CREB-dependent gene expression. Cell. 2012;149:1112–1124. doi: 10.1016/j.cell.2012.03.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Whim MD, Kaczmarek LK. Expression of Kv3.1 in bag cell neurons prevents the induction of a depolarizing afterpotential. J Neurosci. 1998;18:9171–9180. doi: 10.1523/JNEUROSCI.18-22-09171.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. White BH, Kaczmarek LK. Identification of a vesicular pool of calcium channels in the bag cell neurons of Aplysia californica. J Neurosci. 1997;17:1582–1595. doi: 10.1523/JNEUROSCI.17-05-01582.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. White RJ, Reynolds IJ. Mitochondria and Na/Ca exchange buffer glutamate-induced calcium loads in cultured cortical neurons. J Neurosci. 1995;15:1318–1328. doi: 10.1523/JNEUROSCI.15-02-01318.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Wilson GF, Richardson FC, Fisher TE, Olivera BM, Kaczmarek LK. Identification and characterization of a Ca2+-sensitive nonspecific cation channel underlying prolonged repetitive firing in Aplysia neurons. J Neurosci. 1996;16:3661–3671. doi: 10.1523/JNEUROSCI.16-11-03661.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Zenisek D, Matthews G. The role of mitochondria in presynaptic calcium handling at a ribbon synapse. Neuron. 2000;25:229–237. doi: 10.1016/S0896-6273(00)80885-5. [DOI] [PubMed] [Google Scholar]
  101. Zhang S, Yuan JX, Barrett KE, Dong H. Role of Na+/Ca2+ exchange in regulating cytosolic Ca2+ in cultured human pulmonary artery smooth muscle cells. Am J Physiol Cell Physiol. 2004;288:245–252. doi: 10.1152/ajpcell.00411.2004. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Neuroscience are provided here courtesy of Society for Neuroscience

RESOURCES