Skip to main content
The Journal of Neuroscience logoLink to The Journal of Neuroscience
. 2002 Jul 15;22(14):5840–5847. doi: 10.1523/JNEUROSCI.22-14-05840.2002

Presynaptic Mitochondrial Calcium Sequestration Influences Transmission at Mammalian Central Synapses

Brian Billups 1, Ian D Forsythe 1
PMCID: PMC6757942  PMID: 12122046

Abstract

Beyond their role in generating ATP, mitochondria have a high capacity to sequester calcium. The interdependence of these functions and limited access to presynaptic compartments makes it difficult to assess the role of sequestration in synaptic transmission. We addressed this important question using the calyx of Held as a model glutamatergic synapse by combining patch-clamp with a novel mitochondrial imaging method. Presynaptic calcium current, mitochondrial calcium concentration ([Ca2+]mito, measured using rhod-2 or rhod-FF), cytoplasmic calcium concentration ([Ca2+]cyto, measured using fura-FF), and the postsynaptic current were monitored during synaptic transmission. Presynaptic [Ca2+]cytorose to 8.5 ± 1.1 μm and decayed rapidly with a time constant of 45 ± 3 msec; presynaptic [Ca2+]mito also rose rapidly to >5 μm but decayed slowly with a half-time of 1.5 ± 0.4 sec. Mitochondrial depolarization with rotenone and carbonyl cyanidep-trifluoromethoxyphenylhydrazone abolished mitochondrial calcium rises and slowed the removal of [Ca2+]cyto by 239 ± 22%. Using simultaneous presynaptic and postsynaptic patch clamp, combined with presynaptic mitochondrial and cytoplasmic imaging, we investigated the influence of mitochondrial calcium sequestration on transmitter release. Depletion of ATP to maintain mitochondrial membrane potential was blocked with oligomycin, and ATP was provided in the patch pipette. Mitochondrial depolarization raised [Ca2+]cyto and reduced transmitter release after short EPSC trains (100 msec, 200 Hz); this effect was reversed by raising mobile calcium buffering with EGTA. Our results suggest a new role for presynaptic mitochondria in maintaining transmission by accelerating recovery from synaptic depression after periods of moderate activity. Without detectable thapsigargin-sensitive presynaptic calcium stores, we conclude that mitochondria are the major organelle regulating presynaptic calcium at central glutamatergic terminals.

Keywords: mitochondria, calcium imaging, calyx of Held, short-term plasticity, rhod-2, rhod-FF, fura-FF


Neurotransmitter release is triggered by Ca2+ influx through voltage-gated channels (Katz, 1969). Factors regulating the spatiotemporal profile of the presynaptic cytoplasmic calcium transient influence release probability, vesicle recycling, and information processing in the brain. Mitochondria are capable of having an important effect on these processes because they can sequester large quantities of calcium (Duchen, 1999; Nicholls and Budd, 2000), are present at high concentrations in presynaptic terminals, and can be tethered to vesicle release sites (Rowland et al., 2000). Presynaptic buffering of calcium by mitochondria occurs at the mouse (David and Barrett, 2000), lizard (David et al., 1998; David, 1999), and crustacean (Tang and Zucker, 1997) neuromuscular junctions, but is inconsistently observed at the goldfish retinal bipolar cell terminal (Kobayashi and Tachibana, 1995; Zenisek and Matthews, 2000). At the crustacean neuromuscular junction, mitochondrial calcium sequestration over 7–10 min of tetanic stimulation, and its subsequent release, causes post-tetanic potentiation (PTP) (Tang and Zucker, 1997). Although mitochondria are known to buffer calcium and influence secretion of peptides from sympathetic ganglia (Peng, 1998; Cao and Peng, 1999), pituitary gonadotropes (Kaftan et al., 2000), and chromaffin cells (Herrington et al., 1996; Park et al., 1996; Xu et al., 1997; Giovannucci et al., 1999; Sorimachi et al., 1999; Montero et al., 2000), relatively little is known about the role mitochondria play in influencing neurotransmission in the brain because conventional mitochondrial calcium imaging techniques are difficult to applyin situ. Additional complications arise because mitochondrial depolarization blocks both calcium accumulation and ATP synthesis, leading to indirect effects on synaptic transmission attributable to ATP depletion.

We developed a technique to image mitochondrial calcium concentrations ([Ca2+]mito) in single cells and synaptic terminals by including the AM form of a rhod-based dye (rhod-2 or rhod-FF) in the patch pipette. These dyes are cationic (Minta et al., 1989) and are preferentially accumulated into the mitochondria when AM is loaded into cells. However, after AM loading of intact cells with rhod dyes, a substantial cytoplasmic component is still observed (typically 30%) (Kaftan et al., 2000), which must be reduced by up to 24 hr of washing (Simpson and Russell, 1996, 1998), permeabilizing cells with digitonin (Rohacs et al., 1997; Szabadkai et al., 2001), or manganese quenching (Mironov and Richter, 2001). By including the AM dye in the patch pipette we can avoid problems of cytoplasmic loading because cytoplasmic rhod fluorescence is removed by dialysis (back into the patch pipette), leaving fluorescence solely from the mitochondrial compartment. Extensive periods of washing or quenching of cytoplasmic fluorescence were not required, and measurement of cytoplasmic calcium concentrations ([Ca2+]cyto) in the same terminal was possible by also including fura-FF potassium salt in the patch pipette.

In this study, we examined the role of mitochondria in a mammalian glutamatergic synapse known as the calyx of Held (Forsythe, 1994; Borst et al., 1995; von Gersdorff et al., 1997). Using simultaneous electrophysiological recording combined with intracellular calcium imaging of both [Ca2+]cyto and [Ca2+]mito, we examined the influence of mitochondria on neurotransmitter release on millisecond time scales.

MATERIALS AND METHODS

Electrophysiology. Transverse brainstem slices (∼150-μm-thick) were prepared from 10- to 12-d-old Lister Hooded rats, killed by decapitation. Slices were incubated for 1 hr at 37°C in artificial CSF (aCSF), and then stored for up to 8 hr at room temperature before transferring to the experimental chamber for recording at 37°C. Medial nucleus of the trapezoid body (MNTB) neurons and calyces of Held were visualized with infrared differential interference contrast (DIC) optics on a Nikon E600FN microscope with a 60×, numerical aperture 1.0, water-immersion, fluor lens. The aCSF contained (in mm): 125 NaCl, 2.5 KCl, 10 glucose, 1.25 NaH2PO4, 26 NaHCO3, 1 MgCl2, 2 CaCl2, 3 myo-inositol, 0.5 ascorbic acid, 2 Na-pyruvate, 10 TEA-Cl, and 0.001 TTX, pH 7.4, when gassed with 95% O2 and 5% CO2. Whole-cell patch-clamp recordings were made from both the synaptic terminal and postsynaptic cell using thick-walled glass pipettes (GC150F-7.5; Clark Electromedical, Reading, UK) with Axopatch 200B amplifiers (Axon Instruments, Foster City, CA), filtered at 5 kHz (8-pole Bessel filter) and sampled at 10 kHz. Currents were recorded with pClamp software (Axon Instruments). Whole-cell access resistances were <10MΩ for the postsynaptic cell and <30MΩ for the presynaptic terminal, and were compensated >70% with a 10 μsec lag time. The intracellular solution for the postsynaptic cell contained (in mm): 110 CsCl, 40 HEPES, 10 TEA-Cl, 12 Na2-phosphocreatine, and 1 EGTA, pH adjusted to 7.3 with CsOH. The intracellular solution for the presynaptic terminal contained (in mm): 110 CsCl, 40 HEPES, 10 TEA-Cl, 12 Na2-phosphocreatine, 0.2 EGTA, 10 Na-glutamate, 2 Mg-ATP, and 0.5 Na-GTP, pH adjusted to 7.3 with CsOH. A free calcium concentration of 100 nm (measured with fura-2) was attained by the addition of 8 μm CaCl2. A liquid junction potential of 3.4 mV was not corrected for. During paired presynaptic and postsynaptic recordings postsynaptic AMPA receptor responses were pharmacologically isolated, and desensitization and saturation minimized by the addition of 50 μmcyclothiazide and 3 mm kynurenate to the external medium. NMDA receptors were blocked with 50 μmdl-2-amino-5-phosphonopentanoic acid and 10 μm (+)-MK 801 maleate.

Calcium imaging. Calcium-sensitive fluorescent dyes (200 μm fura-FF potassium salt and 10 μm of either rhod-2 AM or rhod-FF AM) were simultaneously added to the presynaptic intracellular solution to visualize the calcium transients in the cytosol and mitochondria of the terminal. After establishment of whole-cell recording, ∼10 min was required for rhod fluorescence to be detected in the mitochondria. Dyes were excited at 380 and 550 nm, respectively, with a Polychrome II monochromator (T.I.L.L. Photonics, Martinsried, Germany). Emitted light was separated by 400 or 575 nm dichroic mirrors and filtered with either a 420 or a 590 nm long-pass emission filter. Fluorescence images were acquired every 10 msec with a PentaMAX cooled CCD camera via a Gen IV image intensifier (Princeton Instruments, Trenton, NJ) and analyzed with MetaFluor software (Universal Imaging, West Chester, PA). Imaging of [Ca2+]cyto and [Ca2+]mito in the same cell was achieved by stimulating once while imaging at 380 nm, followed by once at 550 nm, repeating this alternate excitation and averaging the relevant signals. An interval between each stimulus of 2 min was required to allow re-establishment of baseline calcium levels; hence long duration stable recordings were required for data acquisition. Presynaptic calcium transients were imaged with fura-FF using only 380 nm excitation to enhance the rate of data acquisition and also because kynurenic acid absorbs much of the excitation light below 360nm. The fluorescence was converted to [Ca2+]i using the following equation (Jaffe et al., 1992; Helmchen et al., 1997):

[Ca2+]i=[Ca2+]rest+Kd(ΔF/F)/(ΔF/F)max(1(ΔF/F)/(ΔF/F)max)

A value of −0.76 ± 0.03 was established for (ΔF/F)max at the end of 20 experiments by repeated electrical zapping of the terminal (−1.3 V pulses for 5 msec) to cause a saturating [Ca2+]cyto rise. AKd of 10 μmwas used, as previously determined for fura-FF in this preparation (Bollmann et al., 2000; Schneggenburger and Neher, 2000).

Fluorescent dyes were obtained from Molecular Probes (Eugene, OR), glutamate receptor antagonists from Tocris Cookson (Bristol, UK), TTX from Latoxan (Valence, France), and thapsigargin and Ru360 from Calbiochem (Nottingham, UK). All other chemicals were obtained from Sigma (Poole, UK). Ru360 was diluted from a 1 mm stock made daily in deoxygenated 1 mm Na-ascorbate. Frozen Ru360 stock solution lost activity overnight. Data are expressed as the mean ± SEM. Statistical significance (p < 0.05) was tested with paired or unpaired two-tailed t test as appropriate. All experiments were performed at physiological temperature (35–37°C).

RESULTS

Calcium influx into the calyx of Held (Fig.1A) was triggered by 2 msec step depolarizations under voltage-clamp conditions in the presence of sodium and potassium channel blockers (Helmchen et al., 1997; Borst and Sakmann, 1998). To maintain the balance between physiological calcium homeostatic mechanisms, experiments were conducted at a temperature of 35–37°C, and intracellular mobile calcium buffering was maintained with 200 μmEGTA (Borst and Sakmann, 1996). Inclusion in the patch pipette of the low-affinity calcium indicator fura-FF (Fig. 1B) allowed quantitative imaging of [Ca2+]cyto(Helmchen et al., 1997; Bollmann et al., 2000; Schneggenburger and Neher, 2000) simultaneously with measurement of the presynaptic calcium current generated by P-type calcium channels (Forsythe et al., 1998;Iwasaki and Takahashi, 1998) (Fig. 1C). Presentation of four stimuli at 100 Hz caused a rapid rise of presynaptic [Ca2+]cyto to 8.5 ± 1.1 μm (n = 5) (Fig. 1D) and decayed in a double exponential with time constants (τ) of 43 ± 2 and 311 ± 59 msec (τfast was 83 ± 2% peak amplitude).

Fig. 1.

Fig. 1.

Rapid time course of presynaptic [Ca2+] transients in the calyx of Held.A, DIC image of a calyx of Held (arrow) surrounding a postsynaptic MNTB neuron (asterisk). Scale bar, 15 μm. Patch pipette is to the right.B, Fluorescence image from the same calyx filled with fura-FF via diffusion from the patch pipette. C, Calcium currents (bottom trace) recorded in the calyx of Held in response to four voltage steps (−80 to 0 mV depolarization, 2 msec duration at 100 Hz; top trace). D,[Ca2+]cyto time course during stimulation recorded simultaneously with the calcium current from the same terminal as in C using fura-FF imaging at a frame frequency of 100 Hz (stimulation at arrow as shown in C).

Mitochondrial calcium sequestration was first examined by imaging the fluorescent calcium indicator rhod-2 (Minta et al., 1989), which preferentially accumulates in mitochondria. The AM form of rhod-2 is cationic and membrane-permeant but possesses no detectable calcium-dependent fluorescence. When added via the patch solution, it accumulated in mitochondria because of the large transmembrane potential (Δψm) and was activated by mitochondrial esterases. Once de-esterified, rhod-2 is retained in the mitochondrial matrix, regardless of Δψm. Rhod-2 fluorescence was absent from the cytoplasm and axon (Fig.2C) because of dialysis of cytoplasmic de-esterified rhod-2 (and possibly the esterases themselves) into the patch pipette. The different fluorescent absorbance spectra of rhod-2 and fura-FF permitted measurement of both [Ca2+]cyto and [Ca2+]mito in the same presynaptic terminal. Depolarization-evoked calcium entry into the presynaptic terminal resulted in calcium accumulation in the mitochondria, as indicated by an increase in rhod-2 fluorescence (Fig.2D).

Fig. 2.

Fig. 2.

Slow time course of presynaptic mitochondrial [Ca2+] transients. A, DIC image of the presynaptic terminal. Scale bar, 15 μm. B,Projected stack of fura-FF images focused through the entire terminal.C, Projected stack of rhod-2 images showing that fluorescence was concentrated in the terminal but was absent from the pipette and axon regions. D,[Ca2+]mito rose rapidly on stimulation (as in Fig. 1C, black trace) but decayed slowly. Four groups of four stimuli (the same protocol as in Fig. 1Crepeated 4 times 150 msec apart, gray trace) had a similar time course and amplitude, indicating saturation of the dye. Mitochondrial depolarization (perfusion of 25 μm rotenone and 1 μm FCCP in the presence of 5 μg/ml oligomycin) completely blocked calcium accumulation by the mitochondria after one group of four stimuli (bottom black trace).E, Rhod-FF fluorescence signals were not saturated by four groups of four stimuli (gray trace) compared with one group of stimuli (black trace). Thetraces in D and E are shown normalized to the change in fluorescence after one group of stimuli. F, Summary data show that mitochondrial calcium uptake was inhibited by mitochondrial depolarization and 10 μm TPP+ but not 1 μmthapsigargin. Mitochondrial calcium sequestration saturated rhod-2 (Kd, 570 nm) but not rhod-FF (Kd, 19 μm).Asterisks indicate statistical significance (p < 0.05).

Saturation of rhod-2 by calcium (Kd, 570 nm) severely limited the ability to resolve changes in [Ca2+]mito. Repetition of the stimulus produced no additional increase in fluorescence signal (n = 4; p = 0.27) (Fig. 2D, gray trace), suggesting that the rhod-2 was indeed saturated in these experiments. In contrast, repetitive stimulation of presynaptic terminals after mitochondrial loading with the lower affinity dye rhod-FF (Kd, 19 μm), resulted in progressive summation and greater fluorescent responses to repetitive stimulation (n = 4; p < 0.01) (Fig. 2E, gray trace). Because rhod-FF was not saturated under these conditions, the time course of [Ca2+]mito could be determined. The half rise-time of [Ca2+]mito after four stimuli at 100 Hz was 47 ± 3 msec, and the half decay time 1.5 ± 0.4 sec (n = 3). Although fura-FF and rhod-FF have similar affinities for calcium, the rhod-FF signal was notably slower than the kinetics of [Ca2+]cytorecorded with fura-FF (Fig. 1D), consistent with the notion that these two dyes are signaling calcium changes in different subcellular compartments. The peak [Ca2+]mito could not be estimated from the rhod-FF data because we could not measure a calcium-saturated fluorescent signal from the mitochondria. However, the rhod-2 data (Fig. 2D) represents an increase in [Ca2+]mito of at least 5 μm, taking into account itsKd, saturation, and assuming a resting [Ca2+]mito of 100–200 nm (Babcock et al., 1997; Zhou et al., 1998). The effects of mitochondrial inhibitors on the fluorescence change were used to confirm the mitochondrial origin of the rhod fluorescence signal.

Inhibition of mitochondrial calcium sequestration

Accumulation of calcium by mitochondria was inhibited by three separate methods. First, agents were applied that disrupt Δψm because mitochondrial calcium accumulation is driven by Δψm (Gunter and Pfeiffer, 1990; Gunter and Gunter, 1994; Duchen, 1999; Nicholls and Budd, 2000). Calcium uptake into mitochondria was inhibited by >93% (n = 5; p < 0.05) (Fig.2D) with perfusion of rotenone (a blocker of complex I of the respiratory chain) combined with carbonyl cyanidep-trifluoromethoxyphenylhydrazone (FCCP) (a protonophore) and oligomycin (a blocker of F0F1-ATPase, preventing use of ATP to maintain proton gradients). Second, tetraphenylphosphonium (TPP+), which blocks mitochondrial calcium efflux without effecting ATP production (Aiuchi et al., 1985) (but see Nicholls and Budd, 2000) and mitochondrial calcium uptake (Tang and Zucker, 1997) also blocked the [Ca2+]mito rise after stimulation of the calyx of Held (n = 4;p < 0.05) (Fig. 2F). Third, the inclusion in the patch pipette of Ru360 (a specific inhibitor of the mitochondrial calcium uniporter) (Matlib et al., 1998) prevented the stimulus-induced rise in [Ca2+]mito(n = 3; p < 0.01) (Fig.3F). By contrast, thapsigargin, which blocks calcium pumps into intracellular stores, did not prevent the rise of rhod-2 fluorescence (n = 3;p = 0.19) (Fig. 2F). The antagonist profile outlined in Figure 2 and the prolonged time course of the [Ca2+]mito rise clearly demonstrate that the rhod-AM dyes, loaded by this method, report Δ[Ca2+] from the mitochondrial compartment alone.

Fig. 3.

Fig. 3.

Mitochondrial calcium sequestration buffers [Ca2+]cyto. A, The [Ca2+]cyto transient (using the same stimulation protocol as Fig. 1) was significantly slowed by mitochondrial depolarization, but the peak rise was not significantly altered (97 ± 6% of control; n = 5;p = 0.66). B, The τfast [Ca2+]cyto decay was slowed by mitochondrial depolarization and TPP+but not by thapsigargin (129 ± 9% of control;n = 3; p = 0.08). C, D, Trains of presynaptic depolarization (−80 to 0 mV, 1 msec stimulation repeated 20 times at 200 Hz) produced rises in [Ca2+]cyto that were also significantly slowed by mitochondrial depolarization or 1 μm Ru360. E, F, [Ca2+]mito recorded with rhod-FF increased during stimulation (at arrow as inC). This increase was blocked by mitochondrial depolarization or Ru360. Asterisks indicate statistical significance (p < 0.05).

Mitochondria rapidly buffer presynaptic calcium

Dissipation of Δψm by rotenone and FCCP in the presence of oligomycin significantly slowed the [Ca2+]cytotransient after stimulation of the presynaptic terminal (Fig.3A). The τfast was slowed by 45 ± 6% (n = 5; p < 0.01) (Fig.3B). This slowing was not attributable to ATP depletion from the terminal, because ATP was continually available by dialysis from the patch pipette, and application of oligomycin alone had no significant effect on τfast (110 ± 7% of control; n = 3; p = 0.30). This excludes the possibility that mitochondrial inhibition slows the [Ca2+]cytotransient by reducing cytoplasmic [ATP], which could inhibit the plasma membrane calcium ATPase and, indirectly, Na+-Ca2+exchange. In addition, TPP+ and Ru360, which do not affect mitochondrial ATP production, had similar effects on τfast, slowing it to 203 ± 9% and 412 ± 88%, respectively (Fig.3B,D, note the different stimulus protocols). Thapsigargin had no significant effect on [Ca2+]cyto(n = 3; p = 0.08) (Fig. 3B), consistent with a lack of substantial calcium sequestration into the endoplasmic reticulum. Longer, faster trains of stimuli (1 msec step depolarizations, repeated 20 times at 200 Hz) were also used, which mimic the action potential firing patterns observed at this synapse (Brownell, 1975). As with the briefer stimuli, mitochondrial depolarization significantly slowed the decay of the [Ca2+]cytotransient (Fig. 3C,D), τfast being slowed to 436 ± 32% (from 45 ± 3 to 239 ± 22 msec; n = 6;p < 0.01). Concurrent [Ca2+]mitorecordings with rhod-FF demonstrated that mitochondrial depolarization or inclusion of Ru360 inside the patch pipette prevented mitochondrial calcium sequestration during the longer trains of stimulation (Fig.3E,F).

Mitochondrial depolarization had no effect on the magnitude of the presynaptic calcium current (ICa). A statistically insignificant time-dependent rundown ofICa was observed at physiological temperature (ICa declined by 15 ± 4% over ∼30 min; n = 5; p = 0.06). ICa was stable in experiments performed at room temperature, whereas mitochondrial depolarization still slowed [Ca2+]cytotransients at these lower temperatures (data not shown). TPP+, however, reducedICa by 44 ± 8% (n = 3; p < 0.05), and the peak [Ca2+]cyto by 36 ± 9%, suggesting that TPP+ has additional nonspecific effects on presynaptic voltage-gated calcium channels.

Presynaptic mitochondria influence neurotransmitter release

To assess whether mitochondrial calcium buffering affects neurotransmitter release, simultaneous recordings were made from the presynaptic terminal and postsynaptic neuron (Fig.4A) (Borst et al., 1995; Takahashi et al., 1996). This was done in the presence of cyclothiazide and kynurenic acid to minimize receptor desensitization and saturation (Neher and Sakaba, 2001). A conditioning train of presynaptic step depolarizations (as in Fig. 3C) was followed 500 msec later by one test depolarization to assess the degree of recovery from synaptic depression. EPSCs recorded from the postsynaptic MNTB neuron during the conditioning train showed substantial synaptic depression (Fig. 4B), which recovered within 500 msec. TheICa elicited by the test depolarization was reduced to 90.4 ± 1.6% (n = 3; p = 0.03) compared with the firstICa of the conditioning train (Fig.4C,D) because of the build-up of inactivation of the presynaptic calcium channels (Forsythe et al., 1998). However, the magnitude of this reduction was unchanged by mitochondrial depolarization or inhibiting mitochondrial calcium sequestration with Ru360 (n = 3; p = 0.17 and 0.27, respectively) (Fig. 4D), indicating that the action of these drugs on neurotransmitter release (see below) was not mediated by a change in ICa. The size of the test EPSC was measured under control conditions and after mitochondrial inhibitors in the same terminals. A difference in the test EPSC amplitude should indicate whether mitochondrial calcium sequestration modulates neurotransmitter release. Under control conditions, the test EPSC recovered to 81 ± 4% (n = 3) of the initial train amplitude after 500 msec. After mitochondrial depolarization (reducing Ca2+ sequestration but maintaining presynaptic [ATP] as for Fig. 3) the rate of recovery of the EPSC train was slowed, such that the magnitude of the test EPSC recovered to only 40 ± 5% of control after 500 msec (n = 3; p < 0.01) (Fig.4E,F). Similarly, Ru360 slowed the recovery rate of the EPSC train, resulting in the magnitude of the test EPSC recovering to 51 ± 6% of the initial amplitude (n = 3; p = 0.01). To demonstrate that the synaptic depression induced by mitochondrial depolarization is mediated by changes in the dynamics of the presynaptic calcium transient, the concentration of the calcium buffer (EGTA) in the presynaptic terminal was increased from 200 μmto 1 mm (Borst and Sakmann, 1996; Dittman and Regehr, 1998; Wang and Kaczmarek, 1998; Wu and Borst, 1999). With this additional presynaptic calcium buffering the τfast of the [Ca2+]cytotransient was 35 ± 3 msec under control conditions and only slowed to 73 ± 10 msec during mitochondrial depolarization (compare Figs. 3C, 4G). The effects of mitochondrial depolarization on the rate of recovery of the EPSC from synaptic depression were completely abolished by the additional EGTA (p = 0.38; n = 3) (Fig.4H), strongly implicating [Ca2+]cyto in mediating this effect. Thus, our present data using two independent methods of blocking mitochondrial calcium sequestration, indicate that mitochondria can modulate synaptic transmission by accelerating the recovery from short-term presynaptic depression.

Fig. 4.

Fig. 4.

Presynaptic mitochondrial calcium sequestration influences synaptic transmission. A, The postsynaptic MNTB neuron and presynaptic calyx were simultaneously voltage clamped. Scale bar, 10 μm. B, Trains of stimuli (as in Fig. 3C) produced a markedly depressing EPSC train. At 500 msec after the end of the train, the EPSC had recovered to 81% of control amplitude. C, D,Presynaptic calcium currents recorded in response to the train were unaffected by mitochondrial depolarization. E,Postsynaptic currents normalized to the magnitude of the first EPSC. The absolute magnitude of the EPSC was −4.6 ± 0.4 nA before mitochondrial depolarization and −4.3 ± 0.4 nA after. This 6 ± 2% reduction was a time-dependent run-down and was not statistically significant (n = 3;p = 0.07). Mitochondrial depolarization induced by rotenone and FCCP in the presence of oligomycin had no effect on the initial train amplitude and time course, but it significantly reduced the relative magnitude of the test EPSC. F, Identical results were observed after mitochondrial depolarization or by blocking the calcium uniporter with intracellular Ru360 (1 μm).G, Inclusion of 1 mm EGTA in the presynaptic patch pipette accelerated the decay and reduced the effect of mitochondrial depolarization on the presynaptic calcium transient.H, Presynaptic EGTA at 1 mm abolished the effects of mitochondrial depolarization on the rate of recovery from synaptic depression. Asterisks indicate statistical significance (p < 0.05).

DISCUSSION

We have described a technique for imaging both cytoplasmic and mitochondrial calcium in the same presynaptic terminal at a mammalian central synapse. Using this method we showed that mitochondria rapidly sequester substantial quantities of cytoplasmic calcium, significantly buffering [Ca2+]cyto and influencing neurotransmitter release on millisecond time scales. The selective loading of fluorescent dyes relies on the ability to include the AM ester of the rhod dye in the patch pipette, resulting in accumulation of de-esterified dye in the mitochondria. Loading is thus limited to the mitochondria within the patch-clamped cell, and cytoplasmic fluorescence is removed by dialysis into the pipette. Loading occurs rapidly (within a few minutes), and extensive periods of washing to remove cytoplasmic rhod are not required. Manganese quenching is also not required, which enables simultaneous use of a cytoplasmic dye and avoids problems with manganese block of calcium channels and neurotransmission (Hagiwara and Nakajima, 1966; Llinas et al., 1981; Augustine et al., 1985).

We are confident that the fluorescence labeling seen with the rhod dyes is mitochondrial for three reasons: First, contrasting localization of rhod and fura dyes (Fig. 2B,C). Second, different time course of the mitochondrial calcium signal compared with [Ca2+]cyto(compare Figs. 1D, 2E). And third, the sensitivity of the fluorescence signal to mitochondrial depolarization, TPP+ and RU360. The lack of effect of thapsigargin on the rhod fluorescence signal demonstrates that the signal does not originate from the endoplasmic reticulum and implies that mitochondria are the predominant presynaptic calcium store.

Simultaneous mitochondrial and cytoplasmic imaging

The long excitation wavelength of rhod-2 and rhod-FF (500–600 nm) enables simultaneous measurement of [Ca2+]mito and [Ca2+]cyto by combined loading with shorter wavelength dyes such as fluo-3 (Simpson and Russell, 1998; Boitier et al., 1999; Gonzalez et al., 2000;Trollinger et al., 2000; Collins et al., 2001; Rakhit et al., 2001), fura-2 (Simpson and Russell, 1996; Drummond and Tuft, 1999; Drummond et al., 2000), calcium green (Babcock et al., 1997; Peng and Greenamyre, 1998), and Oregon green BAPTA 5N (David et al., 1998). In our studies we combined rhod AM loading with fura-FF salt in the patch pipette. The salt of the fura dye is not membrane-permeable and will remain in the cytoplasm, ensuring that fura and rhod report [Ca2+] from separate subcellular compartments (Fig. 2B,C). Fura-FF has a relatively low affinity for calcium (Bollmann et al., 2000;Schneggenburger and Neher, 2000) and is ideal for imaging the large presynaptic rises in [Ca2+]cyto with minimal spectral overlap with rhod dyes.

The magnitude of [Ca2+]mito rises

A brief stimulus train (100 msec at 200 Hz) resulted in a rapid rise in [Ca2+]mitothat saturated rhod-2 (Fig. 2D). Rhod-FF is a new low-affinity indicator that overcomes this saturation problem and enables the large rises in [Ca2+]mito to be resolved (Fig. 2E). During stimulation, a minimum [Ca2+]mito rise of ∼5 μm was calculated based on the resting calcium level and saturation of rhod-2. This rise is consistent with studies using mitochondrially targeted aequorin proteins that have shown [Ca2+]mitorises of 5–20 μm in transfected cell lines on agonist application (Rizzuto et al., 1992; Rutter et al., 1996) and ∼300 μm in chronically depolarized bovine chromaffin cells (Montero et al., 2000). Our findings contrast with studies at the lizard neuromuscular junction where [Ca2+]mito rises after stimulation are delayed and do not saturate rhod-2 (David et al., 1998). It seems likely that the proximal location of presynaptic mitochondria to release sites (Rowland et al., 2000) has a major impact on the rate and functional significance of mitochondrial calcium sequestration at the mammalian synapse. The distal location of mitochondria at goldfish bipolar cell synapses may explain their minor contribution to sequestration at that site (Zenisek and Matthews, 2000).

Prevention of ATP depletion

Blocking mitochondrial calcium sequestration had a significant effect on the release of neurotransmitter after a train of stimulation (Fig. 4E). There are four reasons why this effect was not mediated by depletion of ATP after mitochondrial poisoning: first, ATP was continually applied via the patch pipette. Second, oligomycin alone had no effect on EPSC amplitude. Third, TPP+ and RU360 prevented mitochondrial calcium uptake without depolarizing Δψm and depleting ATP (Aiuchi et al., 1985;Matlib et al., 1998). And forth, the effects of mitochondrial inhibitors were reversed by the inclusion of 1 mmEGTA in the intracellular solution, indicating mediation by an intracellular calcium buffering mechanism.

Recovery from synaptic depression

Previous reports have shown that increased presynaptic calcium buffering slows the recovery from synaptic depression (Dittman and Regehr, 1998; Stevens and Wesseling, 1998; Wang and Kaczmarek, 1998). This is contrary to our results where presynaptic terminals with lower calcium buffering (caused by removal of the mitochondrial calcium sequestration) show slower recovery from synaptic depression. The residual calcium hypothesis (Katz, 1969) suggests that neurotransmitter release would be enhanced rather that depressed by this additional free calcium. There is good evidence that [Ca2+]cyto plays multiple roles in regulating the vesicle cycle. For instance, elevated [Ca2+]cyto can inhibit endocytosis at the goldfish bipolar cell synapse (von Gersdorff and Matthews, 1994) and interact with regulators of dynamin to inhibit endocytosis in mammalian nerve terminals (Cousin and Robinson, 2000). Furthermore, the Rab GTPases are activated by Ca2+–calmodulin, causing inhibition of vesicle cycling (Coppola et al., 1999), although Ca2+–calmodulin has also been shown to promote refilling of the rapidly releasing pool of vesicles after high-frequency stimulation (Sakaba and Neher, 2001). The overall effect of [Ca2+]cyto on transmitter release will therefore depend on the balance between these processes. In addition, the subcellular location of the [Ca2+]cyto rise is critical in determining its action. The exogenous mobile calcium buffers applied in previous experiments (Dittman and Regehr, 1998;Stevens and Wesseling, 1998; Wang and Kaczmarek, 1998) will freely diffuse throughout the cytoplasm and influence [Ca2+]cyto close to the site of calcium influx. Because mitochondria are located 100–500 nm from the release sites (Lysakowski et al., 1999; Rowland et al., 2000) and constitute 22% of the presynaptic volume (data from the endbulb of Held; M. Nicol and B. Walmsley, personal communication) (Nichol and Walmsley, 2002), they are ideally situated to sequester calcium from the bulk cytoplasm. Hence, they differentially influence the vesicle cycle compared with mobile exogenous buffers, and they could provide a crucial link between presynaptic metabolic activity and neurotransmitter release.

Functional relevance

Reduced calcium sequestration by mitochondria is associated with aging (Leslie et al., 1985) and neurodegenerative conditions such as amyotrophic lateral sclerosis, Huntington's disease, and Alzheimer's disease (Beal, 2000). In addition, symptoms similar to Parkinson's disease can result from inhibition of mitochondrial respiration by rotenone (Betarbet et al., 2000). Hence, the influence of mitochondrial calcium sequestration on synaptic transmission has important implications for cognitive function. Although there are some studies suggesting that calcium release from endoplasmic reticulum can influence synaptic transmission at other synapses (Llano et al., 2000;Emptage et al., 2001), our data suggest that mitochondria are the major intracellular calcium store in this mammalian glutamatergic terminal. Acting in concert with mobile calcium-binding proteins (Lee et al., 2000) and plasma membrane calcium pumps (Garcia and Strehler, 1999), mitochondria will shape cytoplasmic calcium transients and serve as an intermediary in calcium export. Our observations demonstrate that mitochondria also function in short-term presynaptic modulation on a millisecond time scale and with much lower presynaptic calcium loads than previously thought.

Footnotes

This work was supported by the Wellcome Trust. Thanks to Dr. Euan Brown for assistance in some of the preliminary experiments and to Prof. Michael Duchen for advice on mitochondrial pharmacology and imaging. Thanks also to the Leicester University mechanical and electrical workshops for technical assistance and to Dr. Margaret Barnes-Davies, Dr. David DiGregorio, Dr. Angus Silver, Prof. Nick Standen, and Dr. Martine Hamann for helpful comments on this manuscript.

Correspondence should be addressed to Ian Forsythe, Department of Cell Physiology and Pharmacology, University of Leicester, P.O. Box 138, Leicester LE1 9HN, UK. E-mail: idf@le.ac.uk.

REFERENCES

  • 1.Aiuchi T, Matsunaga M, Nakaya K, Nakamura Y. Effects of probes of membrane potential on metabolism in synaptosomes. Biochim Biophys Acta. 1985;843:20–24. doi: 10.1016/0304-4165(85)90044-3. [DOI] [PubMed] [Google Scholar]
  • 2.Augustine GJ, Charlton MP, Smith SJ. Calcium entry into voltage-clamped presynaptic terminals of squid. J Physiol (Lond) 1985;367:143–162. doi: 10.1113/jphysiol.1985.sp015818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.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]
  • 4.Beal MF. Energetics in the pathogenesis of neurodegenerative diseases. Trends Neurosci. 2000;23:298–304. doi: 10.1016/s0166-2236(00)01584-8. [DOI] [PubMed] [Google Scholar]
  • 5.Betarbet R, Sherer TB, MacKenzie G, Garcia-Osuna M, Panov AV, Greenamyre JT. Chronic systemic pesticide exposure reproduces features of Parkinson's disease. Nat Neurosci. 2000;3:1301–1306. doi: 10.1038/81834. [DOI] [PubMed] [Google Scholar]
  • 6.Boitier E, Rea R, Duchen MR. Mitochondria exert a negative feedback on the propagation of intracellular Ca2+ waves in rat cortical astrocytes. J Cell Biol. 1999;145:795–808. doi: 10.1083/jcb.145.4.795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Bollmann JH, Sakmann B, Borst JG. Calcium sensitivity of glutamate release in a calyx-type terminal. Science. 2000;289:953–957. doi: 10.1126/science.289.5481.953. [DOI] [PubMed] [Google Scholar]
  • 8.Borst JG, Sakmann B. Calcium influx and transmitter release in a fast CNS synapse. Nature. 1996;383:431–434. doi: 10.1038/383431a0. [DOI] [PubMed] [Google Scholar]
  • 9.Borst JG, Sakmann B. Calcium current during a single action potential in a large presynaptic terminal of the rat brainstem. J Physiol (Lond) 1998;506:143–157. doi: 10.1111/j.1469-7793.1998.143bx.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Borst JG, Helmchen F, Sakmann B. Pre- and postsynaptic whole-cell recordings in the medial nucleus of the trapezoid body of the rat. J Physiol (Lond) 1995;489:825–840. doi: 10.1113/jphysiol.1995.sp021095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Brownell WE. Organization of the cat trapezoid body and the discharge characteristics of its fibers. Brain Res. 1975;94:413–433. doi: 10.1016/0006-8993(75)90226-7. [DOI] [PubMed] [Google Scholar]
  • 12.Cao YJ, Peng YY. Caffeine and carbonyl cyanide m-chlorophenylhydrazone increased evoked and spontaneous release of luteinizing hormone-releasing hormone from intact presynaptic terminals. Neuroscience. 1999;92:1511–1521. doi: 10.1016/s0306-4522(99)00077-9. [DOI] [PubMed] [Google Scholar]
  • 13.Collins TJ, Lipp P, Berridge MJ, Bootman MD. Mitochondrial Ca2+ uptake depends on the spatial and temporal profile of cytosolic Ca2+ signals. J Biol Chem. 2001;276:26411–26420. doi: 10.1074/jbc.M101101200. [DOI] [PubMed] [Google Scholar]
  • 14.Coppola T, Perret-Menoud V, Luthi S, Farnsworth CC, Glomset JA, Regazzi R. Disruption of Rab3-calmodulin interaction, but not other effector interactions, prevents Rab3 inhibition of exocytosis. EMBO J. 1999;18:5885–5891. doi: 10.1093/emboj/18.21.5885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Cousin MA, Robinson PJ. Ca2+ influx inhibits dynamin and arrests synaptic vesicle endocytosis at the active zone. J Neurosci. 2000;20:949–957. doi: 10.1523/JNEUROSCI.20-03-00949.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.David G. Mitochondrial clearance of cytosolic Ca2+ in stimulated lizard motor nerve terminals proceeds without progressive elevation of mitochondrial matrix [Ca2+]. J Neurosci. 1999;19:7495–7506. doi: 10.1523/JNEUROSCI.19-17-07495.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.David G, Barrett EF. Stimulation-evoked increases in cytosolic [Ca2+] in mouse motor nerve terminals are limited by mitochondrial uptake and are temperature dependent. J Neurosci. 2000;20:7290–7296. doi: 10.1523/JNEUROSCI.20-19-07290.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.David G, Barrett JN, Barrett EF. Evidence that mitochondria buffer physiological Ca2+ loads in lizard motor nerve terminals. J Physiol (Lond) 1998;509:59–65. doi: 10.1111/j.1469-7793.1998.059bo.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Dittman JS, Regehr WG. Calcium dependence and recovery kinetics of presynaptic depression at the climbing fiber to Purkinje cell synapse. J Neurosci. 1998;18:6147–6162. doi: 10.1523/JNEUROSCI.18-16-06147.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Drummond RM, Tuft RA. Release of Ca2+ from the sarcoplasmic reticulum increases mitochondrial [Ca2+] in rat pulmonary artery smooth muscle cells. J Physiol (Lond) 1999;516:139–147. doi: 10.1111/j.1469-7793.1999.139aa.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Drummond RM, Mix TC, Tuft RA, Walsh JV, Jr, Fay FS. Mitochondrial Ca2+ homeostasis during Ca2+ influx and Ca2+ release in gastric myocytes from Bufo marinus. J Physiol (Lond) 2000;522:375–390. doi: 10.1111/j.1469-7793.2000.t01-2-00375.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Duchen MR. Contributions of mitochondria to animal physiology: from homeostatic sensor to calcium signalling and cell death. J Physiol (Lond) 1999;516:1–17. doi: 10.1111/j.1469-7793.1999.001aa.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Emptage NJ, Reid CA, Fine A. Calcium stores in hippocampal synaptic boutons mediate short-term plasticity, store-operated Ca2+ entry, and spontaneous transmitter release. Neuron. 2001;29:197–208. doi: 10.1016/s0896-6273(01)00190-8. [DOI] [PubMed] [Google Scholar]
  • 24.Forsythe ID. Direct patch recording from identified presynaptic terminals mediating glutamatergic EPSCs in the rat CNS, in vitro. J Physiol (Lond) 1994;479:381–387. doi: 10.1113/jphysiol.1994.sp020303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Forsythe ID, Tsujimoto T, Barnes-Davies M, Cuttle MF, Takahashi T. Inactivation of presynaptic calcium current contributes to synaptic depression at a fast central synapse. Neuron. 1998;20:797–807. doi: 10.1016/s0896-6273(00)81017-x. [DOI] [PubMed] [Google Scholar]
  • 26.Garcia ML, Strehler EE. Plasma membrane calcium ATPases as critical regulators of calcium homeostasis during neuronal cell function. Front Biosci. 1999;4:D869–D882. doi: 10.2741/garcia. [DOI] [PubMed] [Google Scholar]
  • 27.Giovannucci DR, Hlubek MD, Stuenkel EL. Mitochondria regulate the Ca2+–exocytosis relationship of bovine adrenal chromaffin cells. J Neurosci. 1999;19:9261–9270. doi: 10.1523/JNEUROSCI.19-21-09261.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gonzalez A, Schulz I, Schmid A. Agonist-evoked mitochondrial Ca2+ signals in mouse pancreatic acinar cells. J Biol Chem. 2000;275:38680–38686. doi: 10.1074/jbc.M005667200. [DOI] [PubMed] [Google Scholar]
  • 29.Gunter KK, Gunter TE. Transport of calcium by mitochondria. J Bioenerg Biomembr. 1994;26:471–485. doi: 10.1007/BF00762732. [DOI] [PubMed] [Google Scholar]
  • 30.Gunter TE, Pfeiffer DR. Mechanisms by which mitochondria transport calcium. Am J Physiol. 1990;258:C755–C786. doi: 10.1152/ajpcell.1990.258.5.C755. [DOI] [PubMed] [Google Scholar]
  • 31.Hagiwara S, Nakajima S. Differences in Na+ and Ca2+ spikes as examined by application of tetrodotoxin, procaine, and manganese ions. J Gen Physiol. 1966;49:793–806. doi: 10.1085/jgp.49.4.793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Helmchen F, Borst JG, Sakmann B. Calcium dynamics associated with a single action potential in a CNS presynaptic terminal. Biophys J. 1997;72:1458–1471. doi: 10.1016/S0006-3495(97)78792-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.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]
  • 34.Iwasaki S, Takahashi T. Developmental changes in calcium channel types mediating synaptic transmission in rat auditory brainstem. J Physiol (Lond) 1998;509:419–423. doi: 10.1111/j.1469-7793.1998.419bn.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Jaffe DB, Johnston D, Lasser-Ross N, Lisman JE, Miyakawa H, Ross WN. The spread of Na+ spikes determines the pattern of dendritic Ca2+ entry into hippocampal neurons. Nature. 1992;357:244–246. doi: 10.1038/357244a0. [DOI] [PubMed] [Google Scholar]
  • 36.Kaftan EJ, Xu T, Abercrombie RF, Hille B. Mitochondria shape hormonally induced cytoplasmic calcium oscillations and modulate exocytosis. J Biol Chem. 2000;275:25465–25470. doi: 10.1074/jbc.M000903200. [DOI] [PubMed] [Google Scholar]
  • 37.Katz B. The release of neural transmitter substances. Thomas; Springfield, IL: 1969. [Google Scholar]
  • 38.Kobayashi K, Tachibana M. Ca2+ regulation in the presynaptic terminals of goldfish retinal bipolar cells. J Physiol (Lond) 1995;483:79–94. doi: 10.1113/jphysiol.1995.sp020569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lee SH, Schwaller B, Neher E. Kinetics of Ca2+ binding to parvalbumin in bovine chromaffin cells: implications for [Ca2+] transients of neuronal dendrites. J Physiol (Lond) 2000;525:419–432. doi: 10.1111/j.1469-7793.2000.t01-2-00419.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Leslie SW, Chandler LJ, Barr EM, Farrar RP. Reduced calcium uptake by rat brain mitochondria and synaptosomes in response to aging. Brain Res. 1985;329:177–183. doi: 10.1016/0006-8993(85)90523-2. [DOI] [PubMed] [Google Scholar]
  • 41.Llano I, Gonzalez J, Caputo C, Lai FA, Blayney LM, Tan YP, Marty A. Presynaptic calcium stores underlie large-amplitude miniature IPSCs and spontaneous calcium transients. Nat Neurosci. 2000;3:1256–1265. doi: 10.1038/81781. [DOI] [PubMed] [Google Scholar]
  • 42.Llinas R, Steinberg IZ, Walton K. Presynaptic calcium currents in squid giant synapse. Biophys J. 1981;33:289–321. doi: 10.1016/S0006-3495(81)84898-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lysakowski A, Figueras H, Price SD, Peng YY. Dense-cored vesicles, smooth endoplasmic reticulum, and mitochondria are closely associated with non-specialized parts of plasma membrane of nerve terminals: implications for exocytosis and calcium buffering by intraterminal organelles. J Comp Neurol. 1999;403:378–390. doi: 10.1002/(sici)1096-9861(19990118)403:3<378::aid-cne7>3.0.co;2-x. [DOI] [PubMed] [Google Scholar]
  • 44.Matlib MA, Zhou Z, Knight S, Ahmed S, Choi KM, Krause-Bauer J, Phillips R, Altschuld R, Katsube Y, Sperelakis N, Bers DM. Oxygen-bridged dinuclear ruthenium amine complex specifically inhibits Ca2+ uptake into mitochondria in vitro and in situ in single cardiac myocytes. J Biol Chem. 1998;273:10223–10231. doi: 10.1074/jbc.273.17.10223. [DOI] [PubMed] [Google Scholar]
  • 45.Minta A, Kao JP, Tsien RY. Fluorescent indicators for cytosolic calcium based on rhodamine and fluorescein chromophores. J Biol Chem. 1989;264:8171–8178. [PubMed] [Google Scholar]
  • 46.Mironov SL, Richter DW. Oscillations and hypoxic changes of mitochondrial variables in neurons of the brainstem respiratory centre of mice. J Physiol (Lond) 2001;533:227–236. doi: 10.1111/j.1469-7793.2001.0227b.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Montero M, Alonso MT, Carnicero E, Cuchillo-Ibanez I, Albillos A, Garcia AG, Garcia-Sancho J, Alvarez J. Chromaffin-cell stimulation triggers fast millimolar mitochondrial Ca2+ transients that modulate secretion. Nat Cell Biol. 2000;2:57–61. doi: 10.1038/35000001. [DOI] [PubMed] [Google Scholar]
  • 48.Neher E, Sakaba T. Combining deconvolution and noise analysis for the estimation of transmitter release rates at the calyx of Held. J Neurosci. 2001;21:444–461. doi: 10.1523/JNEUROSCI.21-02-00444.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Nichol MJ, Walmsley B. Ultrastructural basis of synaptic transmission between endbulbs of Held and bushy cells in the rat cochlear nucleus. J Physiol (Lond) 2002;539:713–723. doi: 10.1113/jphysiol.2001.012972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Nicholls DG, Budd SL. Mitochondria and neuronal survival. Physiol Rev. 2000;80:315–360. doi: 10.1152/physrev.2000.80.1.315. [DOI] [PubMed] [Google Scholar]
  • 51.Park YB, Herrington J, Babcock DF, Hille B. Ca2+ clearance mechanisms in isolated rat adrenal chromaffin cells. J Physiol (Lond) 1996;492:329–346. doi: 10.1113/jphysiol.1996.sp021312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Peng TI, Greenamyre JT. Privileged access to mitochondria of calcium influx through N-methyl-d-aspartate receptors. Mol Pharmacol. 1998;53:974–980. [PubMed] [Google Scholar]
  • 53.Peng YY. Effects of mitochondrion on calcium transients at intact presynaptic terminals depend on frequency of nerve firing. J Neurophysiol. 1998;80:186–195. doi: 10.1152/jn.1998.80.1.186. [DOI] [PubMed] [Google Scholar]
  • 54.Rakhit RD, Mojet MH, Marber MS, Duchen MR. Mitochondria as targets for nitric oxide-induced protection during simulated ischemia and reoxygenation in isolated neonatal cardiomyocytes. Circulation. 2001;103:2617–2623. doi: 10.1161/01.cir.103.21.2617. [DOI] [PubMed] [Google Scholar]
  • 55.Rizzuto R, Simpson AW, Brini M, Pozzan T. Rapid changes of mitochondrial Ca2+ revealed by specifically targeted recombinant aequorin. Nature. 1992;358:325–327. doi: 10.1038/358325a0. [DOI] [PubMed] [Google Scholar]
  • 56.Rohacs T, Nagy G, Spat A. Cytoplasmic Ca2+ signalling and reduction of mitochondrial pyridine nucleotides in adrenal glomerulosa cells in response to K+, angiotensin II and vasopressin. Biochem J. 1997;322:785–792. doi: 10.1042/bj3220785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rowland KC, Irby NK, Spirou GA. Specialized synapse-associated structures within the calyx of Held. J Neurosci. 2000;20:9135–9144. doi: 10.1523/JNEUROSCI.20-24-09135.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Rutter GA, Burnett P, Rizzuto R, Brini M, Murgia M, Pozzan T, Tavare JM, Denton RM. Subcellular imaging of intramitochondrial Ca2+ with recombinant targeted aequorin: significance for the regulation of pyruvate dehydrogenase activity. Proc Natl Acad Sci USA. 1996;93:5489–5494. doi: 10.1073/pnas.93.11.5489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Sakaba T, Neher E. Calmodulin mediates rapid recruitment of fast-releasing synaptic vesicles at a calyx-type synapse. Neuron. 2001;32:1119–1131. doi: 10.1016/s0896-6273(01)00543-8. [DOI] [PubMed] [Google Scholar]
  • 60.Schneggenburger R, Neher E. Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature. 2000;406:889–893. doi: 10.1038/35022702. [DOI] [PubMed] [Google Scholar]
  • 61.Simpson PB, Russell JT. Mitochondria support inositol 1, 4, 5-trisphosphate-mediated Ca2+ waves in cultured oligodendrocytes. J Biol Chem. 1996;271:33493–33501. doi: 10.1074/jbc.271.52.33493. [DOI] [PubMed] [Google Scholar]
  • 62.Simpson PB, Russell JT. Mitochondrial Ca2+ uptake and release influence metabotropic and ionotropic cytosolic Ca2+ responses in rat oligodendrocyte progenitors. J Physiol (Lond) 1998;508:413–426. doi: 10.1111/j.1469-7793.1998.413bq.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Sorimachi M, Nishimura S, Yamagami K. Sequestration of depolarization-induced Ca2+ loads by mitochondria and Ca2+ efflux via mitochondrial Na+/Ca2+ exchanger in bovine adrenal chromaffin cells. Jpn J Physiol. 1999;49:35–46. doi: 10.2170/jjphysiol.49.35. [DOI] [PubMed] [Google Scholar]
  • 64.Stevens CF, Wesseling JF. Activity-dependent modulation of the rate at which synaptic vesicles become available to undergo exocytosis. Neuron. 1998;21:415–424. doi: 10.1016/s0896-6273(00)80550-4. [DOI] [PubMed] [Google Scholar]
  • 65.Szabadkai G, Pitter JG, Spat A. Cytoplasmic Ca2+ at low submicromolar concentration stimulates mitochondrial metabolism in rat luteal cells. Pflügers Arch. 2001;441:678–685. doi: 10.1007/s004240000466. [DOI] [PubMed] [Google Scholar]
  • 66.Takahashi T, Forsythe ID, Tsujimoto T, Barnes-Davies M, Onodera K. Presynaptic calcium current modulation by a metabotropic glutamate receptor. Science. 1996;274:594–597. doi: 10.1126/science.274.5287.594. [DOI] [PubMed] [Google Scholar]
  • 67.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]
  • 68.Trollinger DR, Cascio WE, Lemasters JJ. Mitochondrial calcium transients in adult rabbit cardiac myocytes: inhibition by ruthenium red and artifacts caused by lysosomal loading of Ca2+-indicating fluorophores. Biophys J. 2000;79:39–50. doi: 10.1016/S0006-3495(00)76272-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.von Gersdorff H, Matthews G. Inhibition of endocytosis by elevated internal calcium in a synaptic terminal. Nature. 1994;370:652–655. doi: 10.1038/370652a0. [DOI] [PubMed] [Google Scholar]
  • 70.von Gersdorff H, Schneggenburger R, Weis S, Neher E. Presynaptic depression at a calyx synapse: the small contribution of metabotropic glutamate receptors. J Neurosci. 1997;17:8137–8146. doi: 10.1523/JNEUROSCI.17-21-08137.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Wang LY, Kaczmarek LK. High-frequency firing helps replenish the readily releasable pool of synaptic vesicles. Nature. 1998;394:384–388. doi: 10.1038/28645. [DOI] [PubMed] [Google Scholar]
  • 72.Wu LG, Borst JG. The reduced release probability of releasable vesicles during recovery from short-term synaptic depression. Neuron. 1999;23:821–832. doi: 10.1016/s0896-6273(01)80039-8. [DOI] [PubMed] [Google Scholar]
  • 73.Xu T, Naraghi M, Kang H, Neher E. Kinetic studies of Ca2+ binding and Ca2+ clearance in the cytosol of adrenal chromaffin cells. Biophys J. 1997;73:532–545. doi: 10.1016/S0006-3495(97)78091-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.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]
  • 75.Zhou Z, Matlib MA, Bers DM. Cytosolic and mitochondrial Ca2+ signals in patch clamped mammalian ventricular myocytes. J Physiol (Lond) 1998;507:379–403. doi: 10.1111/j.1469-7793.1998.379bt.x. [DOI] [PMC free article] [PubMed] [Google Scholar]

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

RESOURCES