Skip to main content
Physiological Reviews logoLink to Physiological Reviews
. 2018 Aug 1;98(4):2063–2096. doi: 10.1152/physrev.00030.2017

Voltage-Gated Calcium Channels: Key Players in Sensory Coding in the Retina and the Inner Ear

Tina Pangrsic 1, Joshua H Singer 1, Alexandra Koschak 1
PMCID: PMC6170976  PMID: 30067155

Abstract

Calcium influx through voltage-gated Ca (CaV) channels is the first step in synaptic transmission. This review concerns CaV channels at ribbon synapses in primary sense organs and their specialization for efficient coding of stimuli in the physical environment. Specifically, we describe molecular, biochemical, and biophysical properties of the CaV channels in sensory receptor cells of the retina, cochlea, and vestibular apparatus, and we consider how such properties might change over the course of development and contribute to synaptic plasticity. We pay particular attention to factors affecting the spatial arrangement of CaV channels at presynaptic, ribbon-type active zones, because the spatial relationship between CaV channels and release sites has been shown to affect synapse function critically in a number of systems. Finally, we review identified synaptopathies affecting sensory systems and arising from dysfunction of L-type, CaV1.3, and CaV1.4 channels or their protein modulatory elements.

I. INTRODUCTION

It has been recognized for some time that exocytosis underlying neurotransmission depends on Ca2+ influx through voltage-gated Ca (CaV) channels (87, 175, 217, 282). More recently, though, it has become apparent that CaV channels also serve as signaling molecules at presynaptic active zones (AZs) (92, 122, 219). Here, we consider the multiple roles played by CaV channels at ribbon synapses (FIGURE 1), the specialized synapses found in auditory and vestibular hair cells, including those in the lateral line organ of teleost fish, in retinal photoreceptors, and bipolar cells.

FIGURE 1.

FIGURE 1.

Presynaptic voltage-gated Ca channels of sensory ribbon synapses. Synaptic transmission at ribbon-type active zones (AZs) is driven by Ca2+ influx through voltage-gated L-type Ca channels, with CaV1.3 being the predominant CaV in the receptor cells of the inner ear [cochlear inner and outer hair cells (IHCs and OHCs; B) and vestibular hair cells (VHCs; A)], and CaV1.4 in synaptic terminals of photoreceptors (C) of the retina. In mammalian and non-mammalian VHCs, additionally CaV3.1 and CaV2 channels were detected, respectively (A). The identity of CaV channels in bipolar cells is unclear (CaV1.X; D), but according to various data all CaV1 isoforms may be present. CaV channels are located in close vicinity to the ribbon, a large synaptic vesicle (SV)-tethering presynaptic structure common to synapses in all depicted cell types.

All of these cells signal in an analog fashion, via graded rather than action potentials; ribbon synapses therefore support continuous exocytosis, generally at high rates (22, 48, 111, 121, 123, 143, 176, 178, 331). Thus ribbon synapses might be designed to speed the delivery of vesicles to release sites (113, 274), although there is also evidence that the ribbon slows vesicle delivery to prevent depletion (159) or unneeded release in inactive retinal synapses (385). Furthermore, ribbons may support molecular priming (335), promote synchronous exocytosis of several vesicles (129), and, perhaps, enable homotypic fusion of vesicles before exocytosis (274, 341).

The ribbon is an organelle composed largely of the vertebrate-specific protein RIBEYE, which was identified and purified first from bovine retinal synaptic ribbons (313). RIBEYE comprises a unique A domain that forms the core of the ribbon by mediating self-aggregation (222, 313, 422) and a B domain that is almost identical to COOH-terminal binding protein 2 (CtBP2; Ref. 313). The B domain also strongly resembles NADH dehydrogenase and is thus likely enzymatically active (319). Proteins interacting with the RIBEYE scaffold are largely conserved between sensory cell types [e.g., piccolino (292), bassoon (84, 178), RIM (RAB3-interacting molecule; Ref. 170), and RIM-binding protein (RIM-BP; Ref. 195) have been found at all ribbon AZs]. Given the unique function of each type of ribbon synapse, the full complement of proteins involved in the vesicle cycle varies between cell types, as we will discuss here.

Exocytosis at all ribbon synapses, though, results from influx of Ca2+ through CaV1-containing Ca channels, which mediate L-type Ca currents (Refs. 15, 25, 95, 225, 280, 349; FIGURE 1). At conventional synapses, in contrast, the presynaptic CaV channels are largely of the CaV2 group (e.g., reviewed in Ref. 55). CaV1 channels are arranged in clusters in the presynaptic density beneath the synaptic ribbon and, depending on the cell type, can be coupled tightly to docked vesicles; herein, we discuss how the physical proximity of CaV channels and vesicles results in so-called “nanodomain” control of exocytosis by highly localized changes in intracellular [Ca2+] (22, 43, 160, 367). Several regulatory molecules that modulate the kinetics of CaV channel gating and (in)activation have been found to influence synaptic transmitter release, and mutations affecting their function may have deleterious consequences for hearing, balance, or vision. In this review, we discuss structural and functional specializations of CaV channel clusters mediating transmitter release at ribbon synapses of sensory receptor cells with the goal of articulating how the first step in synaptic transmission at ribbon synapses permits sensory information to be encoded reliably.

II. PROPERTIES OF VOLTAGE-GATED CALCIUM CHANNELS INVOLVED IN SENSORY CODING AT RIBBON SYNAPSES

A. CaV Channels in Hearing, Balance, and Vision

CaV channels are macromolecular complexes comprising a number of subunit proteins. The core Ca2+-selective channel pore is formed by 1 of 10 α1-subunit isoforms, and accessory subunits contributing to the complex include 1 of 4 different β-subunit isoforms and 1 of 4 different α2δ subunits (FIGURE 2; for review see Refs. 47, 94). The α1 subunit determines the majority of a channel's biophysical and pharmacological properties; accessory subunits modify these properties and in addition influence the abundance and localization of CaV channels, as discussed in sections II and III of this review. CaV channels are grouped according to the characteristics of the currents they mediate: L-type currents arise from channels in the CaV1 family, CaV2 family channels give rise to P/Q-, N-, and R-type currents, and T-type currents are mediated by CaV3 family channels.

FIGURE 2.

FIGURE 2.

Major known CaV1 channel interaction partners at the ribbon synapses of the ear and the eye. The pore-forming CaVα1 subunit composed of four homologous domains (I–IV), each containing six transmembrane segments in yellow with the voltage sensor marked in green. Key interaction partners are depicted with their CaV-related roles indicated in parentheses. Both CaV1.3 and CaV1.4 COOH terminus carry a COOH-terminal modulatory domain (CTM) with interacting proximal and distal COOH-terminal regulatory domains (PCRD and DCRD). The EF-hand motif together with the preIQ and the IQ motif comprise the CDI machinery. Protein kinase A (PKA) phosphorylation of the CaV1.4 COOH terminus is depicted by the red dot. GPI, glycosylphosphatidylinositol anchor of the α2δ subunit; VWA, von Willebrand factor-A domain; AID, alpha interaction domain in the I–II linker region that binds CaVβ subunit; SH3, HOOK, and GK, src-homology-3, HOOK-, guanylate kinase-like-domain of the β subunit; NSCaTE, NH2-terminal spatial Ca2+ transforming element; RIM, RAB3-interacting molecule; C2B and PxxP, C2B domain and proline-rich region of RIM; RIM-BP, RIM-binding protein; CaBPs, Ca2+-binding proteins; CaM, calmodulin; CB, CR, and PVα, mobile intracellular Ca2+ buffers calbindin, calretinin, and parvalbumin α, respectively; GIPC3, GAIP-interacting protein, COOH terminus 3; Bsn, bassoon; ICa, whole cell Ca2+ current; i, single channel current; Po, open probability; VDI and CDI, voltage- and Ca2+-dependent inactivation.

Here, our focus is on CaV1 channels (for review see Ref. 419), which generate the L-type currents and are well suited to mediate neurotransmission evoked by graded potential changes at ribbon synapses. In particular, the gating properties of these channels at ribbon synapses permit rapid responses to fluctuations in membrane voltage elicited by changes in sensory stimulus intensity. The pharmacology of CaV1 channels is well-established, and many clinically relevant compounds, e.g., dihydropyridines (DHPs), are CaV1 modulators.

There are four members of the CaV1 family: CaV1.1, CaV1.2, CaV1.3, and CaV1.4. Both CaV1.3 and CaV1.4 channels activate at relatively hyperpolarized potentials, within the range of graded potential changes in receptor cells, and they inactivate slowly (23, 190, 191, 408). Thus they are highly appropriate for mediating tonic and/or periodic Ca2+ influx necessary for transmission at ribbon synapses.

CaV1.3 channels are expressed most abundantly in inner and outer hair cells of the cochlear (IHCs and OHCs, respectively) and in vestibular hair cells (VHCs; FIGURE 1), but they also are present in several other tissues (IUPHAR target id 530; for review see Ref. 419). The unusual electrophysiological L-type current properties [negative activation threshold, fast activation, and depending on the splice variant (sect. IIB), relatively slow inactivation of Ba and Ca currents], were first described in chick basilar hair cells (424) and later in mouse IHCs (280) and in nonmammalian auditory hair cells (e.g., see Refs. 112, 189, 314). The pharmacological and physiological properties of L-type Ca currents recorded in OHCs are consistent with these channels containing CaV1.3α1 subunits, too (184). Interestingly, CaV1.3 channels expressed in heterologous expression systems or several tissues outside the inner ear (e.g., in the sinoatrial node) show more pronounced inactivation than do channels in hair cells (190, 224, 280, 408), indicating tissue-specific channel specializations and motivating study of alternative splicing, auxiliary subunits, and protein binding partners (see sect. IIB).

While the vast majority of CaV current in IHCs is mediated by CaV1.3 channels (∼90%, Refs. 95, 280), CaV1.3 channels contribute only ∼50% of the total depolarization-evoked Ca current in VHCs in cristae ampullaris of semicircular canals and maculae of otolithic gravity organs (sacculus and utriculus) (20, 95). This might explain why CaV1.3 knockout (KO) animals are deaf but do not suffer from a significant balance impairment (95). The residual whole-cell Ca current in these mice implies that hair cells in both sensory systems may express multiple CaV channels (280). The presence of CaV3.1 channels with somewhat atypical biophysical properties was reported in the mouse inner ear and chicken basilar papilla (202, 203, 260). Due to their low-voltage activation and usually very rapid and strong inactivation, CaV3-mediated Ca2+ influx is responsible for spontaneous activity in neurons and pacemaker cells (for review see Ref. 419). The transient presence of the CaV3.1-mediated Ca current during early hair cell development and/or upon ototoxic drug exposure also suggested its requirement for hair cell maturation and regeneration (202, 203). It has to be noted, however, that other CaV channels also may contribute to maintaining the vestibular synaptic function in CaV1.3 KO animals. Notably, nimodipine-insensitive Ca currents in the hair cells of lower vertebrates were suggested to be mediated via CaV2 channels (230, 304, 350).

Single-channel recordings helped determine the identity of CaV channels in hair cells and showed that CaV1.3 channels display very rapid voltage-dependent activation and deactivation (within <1 ms; e.g., see Refs. 303305, 416418), which enables short delays in synaptic transmission. Together with fluctuation analysis of whole-cell CaV currents (e.g., Refs. 43, 111, 129, 382), they gave insights into the elementary biophysical properties of CaV1.3 and, in the retina, CaV1.4 (366) channels. The reported values vary greatly, depending largely on the experimental conditions (e.g., single-channel conductances for CaV1.3 between 3.5 and 16 pS in hair cells of different species and channels expressed in heterologous expression system; Refs. 39, 129, 304, 416), hampering comparisons among studies. Combining data from recordings in mouse apical IHCs (43, 382, 418) suggests the CaV1.3 single-channel current of approximately -0.14 pA (assuming 1.3 mM extracellular [Ca2+]) and a maximal open probability of 0.2–0.4 (in the absence of BAY K 8644). For CaV1.4, a large discrepancy in the single-channel conductance (3.7 vs. 22 pS) was observed despite similar recording conditions in two studies (i.e., 100 vs. 82 mM Ba2+ as the charge carrier; Refs. 90, 366), which is worthy of further investigation.

Slow inactivation is most pronounced in CaV1.4 channels (23, 191, 236), which are expressed predominantly in the retina, particularly at photoreceptor terminals (FIGURE 1) where they mediate the sustained Ca2+ entry needed for continuous release of neurotransmitters in dark (333, 386). The identity of L-type Ca channels in retina generally and in bipolar cells specifically is controversial. Immunohistochemistry and in situ hybridization suggest the presence of all CaV1 subunits at retinal ribbon synapses (177, 250, 360, 406). Additionally, Ca currents in rod and cone photoreceptors are modulated differently by a variety of neuromodulators, suggesting diversity in CaV channel composition and function (5, 193, 339, 343), but detailed assessment of channel composition is lacking (33, 50, 251, 299).

B. Regulation of CaV1.3 and CaV1.4 Properties

1. Alternative splicing

A key mechanism for regulating the functional properties of CaV1 channels is alternative splicing. Among L-type Ca channels, alternative splicing in the COOH terminus of the CaV1.3α1 subunit is best understood (39, 332, 358). Similar to CaV1.2 channels, full-length CaV1.3 and CaV1.4 channels carry a proximal and a distal COOH-terminal regulatory domain (Refs. 156, 332, 333; FIGURE 2), both of which are putative α-helices that form a noncovalent interaction. This so-called COOH-terminal modulatory domain (CTM) competes with calmodulin (CaM) in binding to the channels’ IQ-domain (216) and thus weakens CaM-mediated Ca2+-dependent inactivation (CDI). It also shifts the voltage dependence of channel activation to more positive voltages and reduces channel open probability (39, 211, 332, 333).

Alternative splicing can affect the presence of one or both components of the CTM or elements in proximity and thereby change channels’ biophysical properties. Alternative splicing of the CaV1.3 COOH terminus generates ‟longˮ (CaV1.3L also termed CaV1.342) and ‟shortˮ channels with different functions in different tissues (332, 358). In cochlear IHCs, containing long and short CaV1.3 splice variants, alternative α1-subunit splicing may contribute to large functional heterogeneity of the channel clusters (Refs. 312, 326, 381, see also Ref. 188 for hair cells from chicken basilar papilla, but see Ref. 262). OHCs preferentially express CaV1.3 channels lacking the IQ motif (CaV1.3IQ∆), which likely explains the almost complete absence of CDI in these cells (326), since the IQ motif acts as the interaction site for CaM (see below).

In human retina, at least 19 splice variants of CaV1.4α1 have been identified (359). Notably, alternative splicing in the IVS3-S4 linker, which may influence S4 voltage sensor function because of its close proximity to the S4 segments (see FIGURE 2 for channel structure), is common in CaV1 channels and likely supports their functional diversity (171). For example, the common CaV1.4Δex32, in which skipping of exon 32 shortens the IVS3-S4 linker, shows altered channel activation (359). Even more abundant is CaV1.4ex16a-, a splice variant with alternative splicing of exon 16. It results in a frameshift and consequently a premature stop codon at IIS6 domain, which was predicted to lead to the formation of a hemi-channel (359). If paired with an alternative splicing at the following exon (17a+1), which in itself also causes a frameshift, the reading frame should be restored and functional channels generated (359). It is unknown whether the splicing of CaV1.4 is cell-type specific, developmentally regulated, and whether it is affected under pathological conditions.

2. CaV channel interacting partners

In numerous ways, CaV1 behaviors are modified via protein-protein interactions (52). Here, we review interactions between CaV1 and the Ca2+-binding proteins CaM and CaBP in sensory receptor cells.

a) calmodulin.

CaM is an important Ca2+ sensor that powerfully regulates CaV channels. The structural and functional highlights of this ‟calmodulationˮ have been reviewed recently (31, 82). Briefly, Ca2+ entering through L-type Ca channels binds to apoCaM preassociated with an interface that includes the COOH-terminal IQ domain (Refs. 29, 99, 108, 179, 356; FIGURE 2). Ca2+-bound CaM then undergoes conformational changes that promote CDI. Ca2+ binding to either lobe of CaM modulates channel function, but the C- and N-lobes show different affinities for Ca2+ (6 times higher for the CaM C- as compared with N-lobe; Ref. 100). Consequently, the two CaM lobes may display different ‟spatial Ca2+ selectivity.ˮ The C-lobe binds Ca2+ entering through the associated channel and releases it very slowly (100), thus displaying ‟local Ca2+ selectivityˮ (356). For the N-lobe, which both binds and releases Ca2+ rapidly (kon of ∼109 to 1010 M−1s−1, koff of ∼104 to 105 s−1; Ref. 100), it has been suggested that brief changes in [Ca2+] cannot shift CaM from the apo state. Therefore, the N-lobe is thought to display a ‟global Ca2+ selectivityˮ for the relatively low changes in [Ca2+] that arise from Ca2+ entry at relatively distant sites (83, 356). Intriguingly, however, the N-lobe was reported to switch its Ca2+ selectivity depending on the presence of the short linear motif, NH2-terminal spatial Ca2+ transforming element (NSCaTE) in the CaV channel (83, 356). NSCaTE, which is only found on the NH2 terminus of CaV1.2 and 1.3 channels (357), binds N-lobe of Ca2+-CaM and endows it with local Ca2+ selectivity. This may be relevant for establishing Ca2+ nanodomains (see sect. III).

CaM also serves as the Ca2+ sensor for Ca2+-dependent facilitation (CDF), but the clear presence of CDF in hair cells has not been demonstrated. There is, however, evidence for paired pulse facilitation of cone Ca currents (379), and mechanisms of voltage-dependent facilitation (VDF) have been studied at photoreceptor synapses (192).

b) calcium binding proteins.

CaM-dependent CDI in L-type Ca channels can be antagonized by calcium binding proteins (CaBPs), which are EF-hand Ca2+-binding proteins with ∼50% homology to CaM (for review, see Ref. 141). They consist of two independent lobes, each containing two EF-hand Ca2+-binding domains, although in CaBPs only two or three are functional (137, 138, 249, 268). Unlike CaM, CaBPs are found primarily in central neurons (CaBP1/caldendrin; Ref. 180), retina (CaBP1, CaBP2, CaBP4, CaBP5; Refs. 136, 138, 299, 334), and the inner ear (CaBP1, CaBP2, CaBP4, CaBP5; Refs. 73, 411).

Although a noncompetitive action of CaBP1 on CaM-dependent CDI has been proposed (265), most models of CaV channel modulation involve a competition between CaBPs and CaM. It has been hypothesized that Ca2+-bound CaBP displaces CaM from the COOH-terminal IQ motif of the α1 subunit (73, 273, 409, 410). For CaBP1, such competitive behavior is well supported by a number of biochemical studies on isolated proteins (106, 107, 399). Despite this evidence, it is unclear how Ca2+-bound CaBPs could displace CaM because CaV channels exhibit a greater affinity for the latter.

This open question is most relevant in the brain, where CaM is far more abundantly expressed than CaBPs (410); the relative concentrations of these proteins in photoreceptors and hair cells are not yet known. One hypothesis is that the competition should occur between the apo-states of the sensors: in the Ca2+-free condition, CaBPs have an advantage over CaM in binding to the CaV IQ domain (107). An alternative, allosteric scheme in which CaM and CaBP4 together bind to CaV1.3 channels also has been proposed (410). Either of the latter mechanisms, though, could permit low concentrations of CaBPs to modulate CaV channels effectively in the presence of higher CaM concentrations. For further discussion on this topic, readers are referred to other recent reviews (e.g., Refs. 30, 31, 105).

c) calcium binding proteins affect sensory neuron function.

In cochlear IHCs, regulation of CaV1.3 channels by CaBPs may be important for synaptic transmission (411). Of the multiple CaBPs detected in IHCs, CaBP2 so far is the only one whose disruption is known to significantly impair hearing in humans (OMIM 607314; Refs. 277, 318; see sect. VI). The lack of CaBP4 affects hair cell physiology only mildly; along with unperturbed hearing in the mouse KO model, this indicates that CaBP4 is only a very modest suppressor of CaV1.3 CDI in IHCs (73).

Milder changes in CaV1.3 CDI can also be observed upon deletion of EF-hand mobile Ca2+ buffers calbindin, calretinin, and parvalbumin-α (270). The mechanisms by which these proteins affect CaV1.3 channels remain elusive, especially as enhanced CDI in mutant IHCs lacking these three buffers could not be reversed upon addition of synthetic Ca2+ chelators. For at least one of Ca2+ buffer proteins, calretinin, a direct interaction with CaV2.1 (but not 1.2) channels has been demonstrated (69), opening the door to future studies.

In the retina, CaBP4 in photoreceptor synaptic terminals interacts with CaV1.4 to regulate Ca2+ influx (136). In addition to reducing channel inactivation, CaBP4 shifts the voltage dependence of CaV1.4 activation to more hyperpolarized potentials and thus increases channel sensitivity to graded membrane potential changes. The effect of CaBP4 on CaV channel activation (but not inactivation) is only observed in the channels containing the CTM (322). Thus it was proposed that CaBP4 forms a collapsed structure around the CaV1.4 IQ motif to disrupt interaction between the IQ domain and the CTM (273). For example, human splice variants CaV1.4ex43* and CaV1.4Δex47 (lacking all or part of the CTM, respectively) exhibit robust CDI and a hyperpolarized activation range in a heterologous expression system (139, 359). The strong CDI (but not a shift in activation range) in CaV1.4Δex47, however, is suppressed in the presence of CaBP4 (139). In photoreceptors, splice variants with the COOH-terminal truncation therefore might show little CDI, similar to full-length channels containing the CTM. Also, the activation range of full-length and truncated CaV1.4Δex47 channels in photoreceptors may be similar due to the counteracting effects of CTM and CaBP4 on the activation range of the full-length channels. This way, both channels with or without CTM would support sustained neurotransmission from photoreceptors in darkness (139). Mutations in CaBP4 (OMIM 608965) cause autosomal recessive congenital stationary night blindness type 2 (CSNB2; also known as incomplete CSNB or iCSNB; Refs. 420, 421; see also sect. VI). This might be explained by the inability of mutant CaBP4 to regulate CaV1.4 channel activation and inactivation effectively (322).

CaBP5 is expressed in some types of retinal bipolar cells, in which it acts to modulate CaV channels, likely CaV1.2 (299). Although CaBP5 KO mice display normal electroretinograms (ERGs) and unperturbed gross retinal morphology, the absolute sensitivity of their rod-mediated ganglion cell light responses is reduced; this indicates that CaBP5 acts rather downstream of photoreceptor synapses (299). CaBP5 also interacts with Munc18–1 and myosin VI; as both are involved in synaptic vesicle cycling and trafficking, it may have additional roles in neurotransmitter release (336). CaBP1 and 2 also modulate synaptic transmission in the inner retina, as ganglion cell responses to light are altered in mice lacking these proteins (334).

In summary, hearing seems to mostly rely on the presence of CaBP2, whereas for vision CaBP4 might be the most relevant. CaV channel function in the sensory organs is likely also modulated by other CaBP members, which however may not be as obvious due to potential functional redundancy among certain protein subfamily members and consequently compensation in the respective single CaBP KO animals.

The function of CaV channels can further be modulated by phosphorylation via various protein kinases. CaV1.4 channels contain a protein kinase A (PKA) phosphorylation site in the distal CTM motif that is involved in the inhibition of the CDI (FIGURE 2). Phosphorylation of this region promotes the occupancy of CaM thus increasing channel open probability and CDI of CaV1.4 full-length channels (311). Indeed, virtually every known neuromodulator affects Ca currents in photoreceptors in various systems (e.g., dopamine, cannabinoids, somatostatin, retinoids; Refs. 5, 162, 339, 343, 380), raising the question of how transmission is maintained over long periods when Ca currents can be affected in so many ways. It has been suggested that PKA-mediated modulation of CaV1.4 channels may contribute to setting the visual sensitivity during the day-night cycle as the PKA activity in retina may also follow circadian rhythm (311). In contrast, no significance of CaV1.3 channel phosphorylation for the function of hair cells has been described yet. Inhibition of PKA activity was shown to have minor effects on CaV1.3-mediated Ca currents in IHCs (397).

3. Modulation of Ca currents by protons released during neurotransmission

Upon synaptic activity, a transient reduction in Ca current is observed at some hair cell, retinal, and electroreceptor ribbon synapses (66, 67, 79, 154, 160, 269, 309, 352, 396). This reduction reflects H+-dependent inhibition of presynaptic CaV channels by the acidic contents of exocytosed vesicles (or by the action of Na+-H+ exchanger, as suggested for the horizontal cell-to-photoreceptor feedback; Ref. 396). Binding of protons to CaV channels partially blocks the channel pore (64, 281) and shields membrane-bound charges, thereby reducing the CaV channel conductance and shifting activation to more positive potentials (183, 194, 281).

Proton-mediated modulation of CaV channels, by limiting excessive Ca2+ influx and subsequent exocytosis, may serve as a mechanism of activity-dependent synaptic plasticity, as suggested for auditory hair cells (66). At the calyceal synapses of the type I VHCs, the released protons may additionally act as nonquantal neurotransmitters (148). Interestingly, H+-mediated block of CaV channels in mammalian IHCs has not been described; this could reflect differences in experimental conditions, the abundance of carbonic anhydrase surrounding hair cells that may oppose excessive acidification of the synaptic cleft, rapid diffusion, rapid clearance of protons away from tiny postsynaptic boutons (66), and/or that IHC synaptic vesicles may be less acidic than others (257).

In the retina, liberation of protons during synaptic transmission could act to match Ca current activation threshold to the photoreceptor resting membrane potential and to set a rate of steady-state exocytosis (21, 79). Acidification of the synaptic cleft by synaptic activity also may be the mechanism of lateral inhibition of rods and cones by horizontal cells (reviewed in Ref. 149). In this scenario, a light stimulus hyperpolarizes horizontal cells, which could drive the influx of protons into the horizontal cells. Alkalinization of synaptic cleft, observed at cone synaptic terminals using a genetically encoded pH indicator (392), would consequently shift the voltage dependence of cone CaV channels towards hyperpolarized potentials, evoking more Ca current at a given membrane potential. Similar mechanisms were suggested for horizontal cell-to-rod feedback (11, 361). In the “reverse operation” of photoreceptors (i.e., they depolarize in darkness and hyperpolarize in light), this would lead to inhibition (that is, affected photoreceptors would respond as if it were darker; Ref. 149). Ca currents also are modulated by other ions in the extracellular environment. These include Zn2+ (7, 403), Cl (362, 364366), and K+ (407). Such modulation might be particularly significant at invaginating photoreceptor synapses, at which diffusion might be restricted and the concentrations of ions might change significantly.

III. CALCIUM CHANNEL SPATIAL ORGANIZATION AS IT RELATES TO EXOCYTOSIS

A. [Ca2+] Microdomain Versus [Ca2+] Nanodomain Control of Exocytosis

The abundance and distribution of CaV channels determines the kinetics and efficiency of exocytosis, the metabolic cost of synaptic activity, and the precision of postsynaptic spike timing. At some synapses the opening of one or at most a few CaV channels triggers the fusion of a nearby synaptic vesicle; such synapses include vestibular and low- to medium-frequency-tuned auditory hair cell ribbon synapses (43, 123, 129, 169, 176), retinal photoreceptor and bipolar cell ribbon synapses (22, 160, 367), as well as some conventional central nervous system (CNS) synapses (reviewed in Ref. 98). In such instances, exocytosis is said to be under control of a “[Ca2+] nanodomain,” a term usually employed when the distance between a CaV channel and a vesicular fusion sensor amounts to less than 100 nm and the opening of a single channel generates a [Ca2+] sufficient to activate the vesicular sensor. In such a scenario, exocytosis of the readily releasable pool of vesicles (RRP) is linearly dependent on the number of open CaV channels. This is in contrast to “[Ca2+] microdomain” control of exocytosis, in which CaV channels are >100 nm from vesicles and Ca2+ influx pooled from multiple channels is necessary to elevate [Ca2+] sufficiently to elicit exocytosis (42; reviewed in Ref. 98). In the case of “microdomain” control, exocytosis shows high cooperativity and highly nonlinear dependence on the number of open channels.

Microdomain control of exocytosis reduces jitter and noise introduced by stochastic channel openings and therefore enhances the “signal-to-noise” ratio of AP-dependent signaling (232). As well, by permitting multiple channels to shape [Ca2+] at release sites, microdomain control has the potential to provide a mechanism by which CaV channel modulation regulates Ca2+-dependent short-term plasticity of transmitter release (317). Nanodomain control of exocytosis, on the other hand, provides increased speed and efficacy of synaptic transmission and greater sensitivity and dynamic range of the response (reviewed in Refs. 98, 232). A [Ca2+] nanodomain might further be advantageous for synchronizing multivesicular release (129, 160, but see Refs. 238, 252, 331). Therefore, it is ideal for encoding graded sensory signals with a large dynamic range in the retina and cochlea.

B. Topology of CaV Channels and Synaptic Vesicles at the (Inner) Hair Cell Ribbon-Containing AZ

The spatio-temporal profile of an intracellular [Ca2+] increase around a single CaV channel depends on its conductance, the concentration and binding kinetics of the intracellular Ca2+ buffers, and the potential spatial constraints (e.g., diffusion barriers; Ref. 129) (232, 254, 317). As the intrinsic Ca2+ dependence of exocytosis is highly nonlinear, with Ca2+ sensors of exocytosis requiring a high degree of cooperativity of bound Ca2+ to trigger transmitter release (35, 40, 41, 143, 145, 316, 402), the release of a synaptic vesicle strongly depends on its distance to the CaV channel.

Direct anatomical study of the topology of CaV channels and synaptic vesicles is difficult and initial estimates of distances between channels and release sites often come from analysis of the effects of the slow Ca2+ chelator, EGTA, on exocytosis (255). Consistent with nanodomain coupling, EGTA has small effects on transmission from auditory hair cells in a number of species [mice (34, 169), frogs (169), turtles (315)]. In the apical IHCs after the onset of hearing, the estimated average distance between a CaV channel and a release site is 15–17 nm (270; in line with Refs. 43, 123, 402), which is essentially half of a vesicle diameter and thus likely the limit to how close a vesicle can be placed next to a CaV channel.

This estimate is supported by a number of imaging studies. Serial section electron micrographic (EM) analysis of the mature apical IHC synapse demonstrates presynaptic densities occupied largely by the presynaptic ribbon (∼97%) and covering ∼420 × 80 nm2 (111, 402). On average, one ribbon tethers 70 vesicles (taking into account that the tomogram typically contains a half of the ribbon) (170, 402; FIGURE 3); 14 of these are found near the center of the AZ, 20–50 nm from the plasma membrane, and likely constitute the readily releasable pool, or RRP (170, 344, 384). Roughly 40% of the RRP appears to be tethered to the plasma membrane, with a tether length of ∼20–25 nm (170, 383). RRP vesicles can be released by brief (<20 ms) depolarizations to potentials eliciting maximal currents (with 2–10 mM extracellular [Ca2+]; Refs. 34, 73, 253, 271).

FIGURE 3.

FIGURE 3.

Active zone (AZ) models of ribbon synapses. A–H: cochlear inner hair cell (IHC) ribbon synapses after the onset of hearing: a section of an EM tomogram (A, E) and 3D model reconstruction (B and C, F and G). C and G: view on the AZs from the top. Ribbon and ribbon-associated vesicles were omitted for clarity. D and H: illustrations depicting the three reconstructed AZs and showing a possible arrangement of CaV channels. Note that the EM section typically contains a half of the ribbon, consequently the reconstructions in F and G lack a part of an AZ. All scale bars: 100 nm. [A–C modified from Strenzke et al. (344). E–G courtesy of Rituparna Chakrabarti.] I and J: EM analysis of rod bipolar (RB) terminals. I: transmission electron micrographic image of a RB terminal illustrates two ribbons (arrows) at dyad synapses, presynaptic to AII and A17 amacrine cells. Scale bar: 500 nm. J: illustration of an AII dendrite contacting a single RB terminal varicosity. Note that the AII is postsynaptic to vesicles on ribbons as well as vesicles near (<24 nm) the plasma membrane but not on ribbons. Illustration is based on EM reconstruction of a retinal volume imaged by scanning block-face electron microscopy (46, 237). Scale bar: 1 µm. [J modified from Mehta et al. (237); I is unpublished data acquired by J. H. Singer.]

CaV channels are clustered below each ribbon, as shown by wide-field, confocal, STED, and electron microscopy (43, 111, 158, 258, 302, 374, 402). These clusters appear as one to three stripes flanked by the presynaptic scaffold protein bassoon (111) and are colocalized with ribbons, such that each ribbon may be decorated by a single elongated CaV channel cluster (258). Estimates of the number of channels in a hair cell come from electrophysiological measurements of macroscopic and single-channel currents, and from these, the total number of CaV channels in a hair cell is estimated to be 1,000–3,000 depending on species (43, 111, 129, 302, 382, 416). How much of the macroscopic Ca current is carried through channels located outside synapses is not entirely clear. But, it appears that in the IHCs after the onset of hearing, few or no CaV channels may be extrasynaptic (43, 402), so dividing the number of channels by the number of ribbons yields an estimate of 40–180 channels per AZ (see TABLE 1).

Table 1.

Estimated average number of CaV channels per an AZ of hair cells in different organs and species

Hair Cell Type, Species Whole Cell No. of Channels No. of Synapses/Ribbons per Hair Cell Approx. Average No. of Channels per an AZ
Mature mouse IHC 1,450–1,850 (43, 111, 382) 12–16 (2–3 wk old apical-mid region) (43, 111, 382) 80–120
Gerbil IHC 2,800 (416) 14–22 (basal adult IHC) (167, 416) 115–180
P5–7 mouse type I VHC 1,175 (382) ~8 (96, 382) 160
Frog saccular and papilla hair cell 1,800–~3,000 (129, 302) 19 (302), 55 (129) 90–100, 40–50

It was assumed that there are a few (500; Ref. 43, 10%; Ref. 416) or no extrasynaptic channels. IHC, cochlear inner hair cell; VHC, vestibular hair cell; AZ, active zone.

Mathematical modeling has revealed possible IHC AZ topographical arrangements (402). A model with 36 randomly distributed as well as a few “private” CaV channels (i.e., channels in direct contact with Ca2+ sensors on synaptic vesicles, thus molecularly coupled) reproduced well the estimated apparent Ca2+ cooperativity of IHC RRP exocytosis at smaller AZs. Such arrangement might represent the realistic mature IHC AZ topography (270, 402). For larger AZs, it seems like an additional spatial constraint around the “private” channels (i.e., an “exclusion zone”) has to be introduced to keep the control of exocytosis under experimentally observed nanodomain regime (Supplemental Figure 9 from Ref. 402).

It is worth considering the molecular mechanisms that might promote nanodomain coupling by regulating the distance between synaptic vesicles and CaV channels. In the immature calyx of Held, for example, it was suggested that septin 5, a filamentous GTP/GDP binding protein, may form the physical barrier preventing vesicle docking to the AZ, thus also maintaining a microdomain control of exocytosis (412). Upon development, septin molecules reportedly undergo a spatial reorganization, being less frequently detected at the mature AZs, thus allowing a tighter docking of synaptic vesicles to the AZ and the movement of CaV channels closer to release sites. At the IHC synapse, a developmental tightening of the vesicle-channel coupling also has been observed to accompany maturational changes in AZ arrangement (Ref. 402; see also sects. IIID3 and V), but a potential role for septin or a protein with similar function in determining the coupling distance has not been addressed. Septin itself might not be found at IHC synapses, which appear to lack neuronal syntaxins (261), the binding partners of septin at conventional synapses (27).

Future work using high-resolution microscopy or SDS replica labeling might be able to validate models of AZ organization, including the distribution of individual proteins necessary to maintain nanodomain coupling. Indeed, a recent study incorporating confocal microscopy, optical fluctuation analysis, and 3D-STED nanoscopy has provided the most accurate measurements of single mouse IHC AZs to date: 30–360 CaV channels are found at each AZ (of an apical IHC after the onset of hearing), far greater variability than suspected previously (258). Another issue to address is the anatomical basis of observed nanodomain [Ca2+] heterogeneity (110). It is interesting to consider how AZ shape and size and tonotopic variability in CaV channel properties (e.g., channel inactivation, open probability; Refs. 168, 244, 417) could affect signal coding at individual synapses. For example, CaV channel-vesicle coupling in IHCs may be looser at the base of the cochlea, such that exocytosis there is under [Ca2+] microdomain control, a difference also observed between high- and low-frequency-tuned auditory bullfrog hair cells (169). It was proposed that microdomain coupling may be more suited for the high-frequency-tuned IHCs of the basal cochlea, which cannot follow the frequency components of the sound, to accurately encode timing and intensity, whereas nanodomain coupling of the low-to-medium frequency-tuned IHCs would enable accurate phase-locking to stimuli in the mid- to apical cochlea (169). Tonotopic differences were also observed in the elementary properties of IHC CaV1.3 channels in the gerbil IHCs (416, 417) and, to a lesser degree, mouse IHCs (244). Reported differences in open probability may be due to tonotopical differences in channel modulation or contributing CaV1.3 splice variants (39, 417) and suggest that CaV channel properties along the cochlear tonotopic axis may be optimized to best suit the sound coding requirements (i.e., phase-locking to low-frequency sounds towards the cochlear apex; Ref. 417).

Understanding the organization that gives rise to nanodomain coupling has the potential to solve one of the mysteries of ribbon synapse physiology generally and hair cell synapse function specifically. At mammalian IHC synapses, two quite distinct types of postsynaptic events are observed: very large monophasic excitatory postsynaptic currents (EPSCs) and small multiphasic EPSCs (58, 121, 123, 126, 308). The two were suggested to reflect highly and poorly synchronized multivesicular release, respectively (121). At frog hair cell synapses, which operate at lower frequencies than mammalian IHCs, EPSCs are far more uniform (129, 176, 204).

One of the proposed mechanisms for multivesicular release in hair cells is the synchronization of release sites (126, 129). Considering the absence of a synchronizing electrical signal like an action potential (AP), and given the presence of [Ca2+] nanodomain control of exocytosis, it is difficult to envision how the graded receptor potential of an IHC could synchronize the simultaneous opening of several CaV channels to synchronize fusion of up to six vesicles. Indeed, the mammalian IHC AZ topology as currently understood does not seem to permit such synchronization by CaV channels (258, 402). According to some studies, the resting open probability of CaV1.3 may be too low for a number of them opening simultaneously to trigger synchronized vesicle fusion in the absence of stimulation (129, 416). Additionally, several lines of experimental data and theoretical analysis do not support the presence of release site coordination at the mammalian hair cell synapse: 1) a disproportion of adapted spiral ganglion neuron (SGN) sound-driven spike rates as measured in vivo (i.e., 200–400 Hz) and as predicted from presynaptic sustained vesicle release rates [maximally 120 Hz, obtained when considering a maximal sustained release rate of 700 Hz (271), a fusion of 6 vesicles per one EPSC (126) and each triggering one spike (308)], 2) the apparent lack of an mEPSC and the inability to reconstruct multiphasic release events from combining several desynchronized theoretical mEPSCs, and 3) the persistence of EPSC amplitude and shape heterogeneity in the near absence of extracellular Ca2+ or upon IHC hyperpolarization, which should both generate conditions preventing synchronization of events by Ca2+ (58).

Other explanations for coordinated multivesicular release from mammalian IHCs are equally unsatisfactory: there is little experimental evidence for homotypic fusion before a single release event, and homotypic fusion seems incompatible with the theoretical considerations based on available experimental findings (58). Experimental exocytic quantal size distributions could only be well reproduced in a model where homotypic fusion rate decreased not only with the distance from the AZ but also with the size of the fused vesicles. A mechanism that would prevent large vesicles from further homotypic fusion (i.e., homotypic fusion “ad perpetuum”) and enable fusion with the plasma membrane, however, is not known, and expected vesicle size distribution in such case shows a strong mismatch to the actual EM data (58). Thus it has been argued that what appears to be multivesicular release is actually quantal release (58). Experimental data in favor of that hypothesis include the finding of 1) comparable EPSC charges observed among EPSC of different shapes and of 2) big clusters of AMPA receptors, which according to the mathematical modeling should be able to support large EPSC events (58).

At frog hair cell synapses, on the other hand, evidence for quantal and multiquantal transmission is quite strong (e.g., the presence of mEPSCs and disappearance of large EPSCs upon hair cell hyperpolarization) (204). Modeling studies have suggested that a single CaV channel opening could generate a large enough change in [Ca2+] in the restricted space below the ribbons to synchronize vesicle fusion and evoke multivesicular release (129). This is an interesting hypothesis, which needs the following considerations: 1) to what degree a ribbon can act as a diffusional barrier for Ca2+ needs to be further addressed (Ca2+ imaging experiments show no exclusion of Ca2+ from the ribbon volume, at least not in mammalian IHCs upon longer stimulation; Refs. 110, 262, 402), and 2) at conventional synapses (81, 157) temporal jitter in the process of exocytosis occurs downstream of Ca2+ entry and binding to molecular sensors. Conflicting observations made at mammalian versus frog hair cell ribbon synapse as well as retinal ribbon synapses raise a general, and important, question of whether there is a prototypical ribbon synapse.

C. CaV Channel-Vesicle Coupling at Retinal Ribbon Synapses

Like hair cell synapses, retinal ribbon synapses are capable of exocytosis on a very rapid time scale: several vesicles/AZ/ms (22, 142, 143, 331). As in hair cells, exocytosis from retinal ribbon synapses studied in slice preparations can be evoked by the opening of as few as one channel per AZ (22, 160), whereas exocytosis from isolated neurons—goldfish bipolar cells in particular—appears to require Ca2+ influx through multiple CaV channels (70). This could arise from species differences, or it could reflect alterations in AZ geometry arising from disruption of extensive interactions between presynaptic proteins (e.g., CaV channels) and elements in the extracellular matrix and on postsynaptic cell(s) (28). Additionally, coordinated multivesicular release is observed at retinal ribbon synapses, particularly mammalian rod bipolar cell synapses, at which it closely resembles coordinated multivesicular release at frog hair cell synapses (160, 205, 238, 331).

In the absence of diffusion barriers that confine free Ca2+ (e.g., Refs. 22, 129), depolarization-evoked increases in [Ca2+] within a nanodomain are relatively small, implying that exocytosis from intact retinal ribbon synapses is mediated by vesicular sensors with relatively high affinity for Ca2+ (effective KD in the 1–10 µM range; Refs. 97, 160). Indeed, direct measurements of sensor KD at photoreceptor AZs support this assertion (301, 367), although sensor KD in isolated bipolar cells appears to be far higher (∼200 µM; Ref. 143). Thus activity-dependent alterations in CaV channel distribution that occur as membrane is redistributed during the cycle of exo- and endocytosis have the potential to affect the efficiency of CaV channel-release site coupling (240, 241).

The Ca2+ sensor mediating fast modes of exocytosis from retinal ribbon synapses is likely synaptotagmin 1 or 2 (144, 376), and the efficiency of the release process at these synapses appears to be enhanced by the action of complexin 3, a member of the complexin family of proteins, which facilitate and synchronize evoked neurotransmitter release by stabilizing partially assembled SNARE complexes (252, 369). Indeed, fast, nanodomain-evoked exocytosis is impaired significantly at bipolar cell synapses lacking complexin 3 (252, 378). Complexin 3 does not affect the distribution of vesicles at the AZ; rather, it appears to modulate the efficiency of the release machinery itself (252). Complexin 4 also can function at photoreceptor synapses, but its expression and significance seem to be limited to circumstances in which complexin 3 is absent (10, 294).

It is important to note that not all transmission from retinal ribbon synapses is driven by [Ca2+] nanodomains. Indeed, at the same synapses that exhibit fast, nanodomain-evoked release, sustained presynaptic depolarization generates sufficient Ca2+ influx to saturate endogenous buffers and elevate global [Ca2+] significantly, in some instances with the assistance of Ca2+ released from intracellular stores (CICR), to evoke tonic exocytosis (9, 51, 6062, 237). Slower modes of exocytosis do not require complexin 3, indicating that [Ca2+] is elevated sufficiently to activate release machinery operating in low efficiency modes (252). Release driven by elevated global [Ca2+] seems to be especially relevant to rods, in which CICR-mediated neurotransmitter release from sites away from synaptic ribbons might underlie more than half of the sustained postsynaptic response (60); this release mode could promote tonic release under dim light conditions.

D. Regulation of CaV Channel Membrane Expression

The abundance of CaV channels at the plasma membrane depends on several factors that influence their trafficking, localization, and degradation. The cellular mechanisms that underlie these processes are only partially known. In this review we will highlight some of the mechanisms involved.

1. Trafficking of CaV channels to the plasma membrane

Classically, CaV trafficking to the plasma membrane is believed to depend on the presence of auxiliary β (36) and α2δ subunits (54; FIGURE 2). The β subunit may act to promote protein trafficking out of the endoplasmatic reticulum (ER) by masking a retention signal residing in the I–II loop of the α1 subunit (CaV2.1; Ref. 36) or in the COOH-terminal region (CaV1.2 and CaV2.2; Ref. 6). Alternatively or additionally, the β subunit may guide protein folding, prevent ubiquitination and degradation of CaVα1 subunits (6, 267, 388), and/or mediate weaker secondary interactions with other channel regions (6; reviewed in Refs. 47, 93, 94). The presence of a putative ER retention signal in the COOH-terminal region of the α1 subunit also suggests that CaM may regulate CaV trafficking (390).

For the α2δ subunit, its metal ion-dependent adhesion site (MIDAS) motif within the von Willebrand factor-A (VWA) domain might act as a chaperone to prevent unfolding and/or facilitate channel trafficking towards the plasma membrane (32, 54, 115). Alternatively, an intact MIDAS motif could be essential for proper coupling of CaV channels and exocytosis by increasing channel density near AZ release sites (153). Further molecules that may affect trafficking of CaV channels are CaM kinase (329) and GIPC3 (GAIP-interacting protein, COOH terminus 3). The latter was suggested to contribute to asymmetrical abundance of CaV channels in the AZs along the modiolar-pillar axis of each IHC (262; see sect. IV).

Some IHCs express multiple β subunits (β2, β3 and β4) (197), but β2 (likely β2a) is the dominant isoform (256). Whereas deletion of β3 or mutation of β4 subunit leaves hearing unperturbed and shows minimal effects on CaV channel abundance in the IHC plasma membrane (197), the absence of β2 severely affects hearing (256). Animals lacking β2 show strongly reduced IHC synaptic Ca2+ influx and exocytosis, attributable to a ∼60% reduction in plasma membrane CaV channel expression (256). This in turn is most likely due to defective channel trafficking to the plasma membrane (256). β2-Deficient IHCs further display a lag in development, supporting the notion of the importance of sufficient prehearing Ca spiking for proper IHC maturation (see sect. V). For the function of OHCs, β3 may be of greater importance (197).

In contrast to β subunits, it is less clear which isoforms of α2δ are present in cochlear hair cells. It was first suggested that α2δ3 may be required for IHC function because its absence leads to impaired acoustic startle reflex in mice. But, detailed investigation of the α2δ3 KO mouse model revealed that hearing impairment is related to dysfunction of presynaptic SGN signaling in the cochlear nucleus rather than a defect in hair cell function (279). In fact, IHCs showed no presence of α2δ3 mRNA in the adult stage and very low levels before the onset of hearing. Recently, the cochlear function of a mutant mouse line (ducky mice) with truncated α2δ2 subunit was investigated and revealed decreased Ca currents and voltage sensitivity in IHCs, suggesting that this is the predominant α2δ subunit of the IHC CaV1.3 (101). Those mice also display defects in cochlear amplification, which together with affected Ca currents likely explains an observed mild hearing impairment.

In the retina, all β isoforms are expressed, and individual isoforms are partially segregated in different layers (17, 185). β2 seems to be the predominant accessory subunit (185), and especially the outer plexiform layer (OPL) is strongly stained by the antibodies against β2 (17). The staining pattern resembles the immunostaining of synaptic ribbon markers, implying this is the β subunit of photoreceptor terminals (17), which also contain CaV1.4 (see above). Recently, a novel, retina-specific β2 splice variant, named β2X13, was identified. It was shown that CaV1.4 properties are differently modulated by the two splice variants, β2X13 and β2a (199), which vary in their HOOK domain (see FIGURE 2) containing key determinants controlling CaV channel inactivation (298). Both splice variants were detected in photoreceptor terminals, suggesting the presence of functionally heterogeneous CaV1.4 channels. Intriguingly, the labeling pattern of CaV1.4 channels was altered in retinas of the mice lacking β2, pointing to the importance of the β2 subunit for proper retinal targeting of the CaV1.4α1 subunits. Furthermore, the phenotype of these mice was in many aspects similar to that seen in CSNB2 patients (see sect. VIB), supporting the role of this subunit in retinal synaptic transmission; in contrast, CNS-β1-, β3-, and β4-deficient mice show normal ERGs (18). In heterologous expression systems, coexpression of CaV1.4α1 and α2δ1 with β2 (as compared with β3) subunit significantly impacted VDI (191) such that CaV1.4 current even better resembled the L-type Ca currents of the native photoreceptors.

Several lines of evidence support the idea that CaV1.4 channels form a complex together with β2 and α2δ4 accessory subunits. Similarly to β2, antibodies directed against α2δ4 label the OPL of the retina (75), and α2δ4 was shown to colocalize with CaV1.4α1 and β2 subunits at structures resembling photoreceptor synaptic ribbons (199). Current densities of CaV1.4 channels heterologously expressed with either α2δ4 or α2δ1 subunits were comparable, a finding that was independent from the coexpression with β2 subunit (12, 199), suggesting that regions of the α2δ protein involved in CaV1.4 membrane targeting are largely conserved between α2δ4 and α2δ1. A spontaneous frameshift mutation in mice resulted in a truncation of the α2δ4 protein, which was expressed at very low levels (404). Corresponding to the defective retinal signal transmission (loss of ERG b-wave), the photoreceptor synaptic layer of the mutant mice is significantly thinner. This may be explained by a recent observation in rods, demonstrating that α2δ4 is essential for establishing synaptic contacts (393). In humans, two mutations in exon 19 and 25 of the CACNA2d4 gene that encodes α2δ subunits have been associated with slowly progressive cone dystrophy (8, 405). Interestingly, inclusion of an alternative exon 25b was recently reported to result in an α2δ4 isoform that completely lacks the δ peptide, holding the α2δ subunit at the membrane (12). Lack of the membrane anchor should result in secretion of the remaining part of the protein, which could exert different, also non CaV channel related targeting functions in the CNS (for review see Ref. 94). Immunohistochemistry revealed α2δ3 immunoreactivity in cell bodies of the ganglion cell layer and the inner nuclear layer (INL) as well as the inner plexiform layer (IPL) and OPL processes. The expression of α2δ3 in multiple cell types including ganglion, amacrine, and bipolar cells and photoreceptors, but not horizontal cells, suggests that this subunit might widely participate in retinal CaV channel signaling (276). So far however, no functional data from α2δ3-deficient animals supporting this hypothesis have been reported.

Whereas β2 acts as the β subunit of CaV1.4 channels, other β isoforms in the retina may pair with other CaV subtypes. The IPL is occupied mostly by β3 and β4 subunits; expression of β2 in this area is diffuse, but two distinct narrow bands stained by antibodies against β3 are seen, suggesting its presence mainly on the cell bodies of cholinergic amacrine cells and in ON and OFF synaptic layers of the IPL. Rod ON bipolar cells might express β2 and/or β1. Finally, β1 subunit seems to be expressed predominantly in glial Müller cells, as it was found on cell bodies in the INL and processes within the inner and outer limiting membrane (17).

2. Targeting of CaV channels to the AZs and stabilization of the channels at the plasma membrane

Little is known about the mechanisms by which CaV1.3 and 1.4 are targeted to synapses. In contrast to the CaV2 channels that mediate synaptic transmission at conventional synapses, CaV1 channels lack the synaptic protein interaction (synprint) motifs (in the II-III loop), MINT1, and CASK interaction sites that are involved in targeting CaV channels to presynaptic AZs (234, 247, 297, 327, 328). Thus CaV channel interactions with other presynaptic proteins might be particularly important for ribbon synapse function.

In addition to their roles in trafficking, auxiliary CaV channel subunits also stabilize L-type and most non-L-type CaV channels in the plasma membrane. The α2δ subunits appear to reduce the turnover of CaV channels (32, 115). The β2a subunit, which is palmitoylated and therefore stabilized in the membrane, may even anchor channels to the plasma membrane (65). A central role in stabilizing CaV channels close to the AZ, however, is ascribed to the RIMs. These presynaptic scaffold proteins are believed to hold channels at the presynaptic membrane either by directly binding to the CaVα1 subunit or indirectly by binding to RIM-BP (4) and/or β subunits (reviewed in Ref. 351; FIGURES 2 and 4).

FIGURE 4.

FIGURE 4.

CaV channel-tethering complex at the ribbon active zone (AZ). L-type Ca channels at the ribbon synapses likely bind RIM via RIM-BP and a palmitoylated auxiliary β2 subunit. RIM and RIM-BP may bind AZ proteins (potentially CAST/ELKS and/or others) to hold CaV channels below the ribbon (presumable interactions of RIMs are depicted by dotted arrow). A Bassoon and/or RIBEYE were further suggested to organize presynaptic CaV channel clusters. Whereas CaV1.3 may bind RIM also directly, the molecular interaction between CaV1.4 and RIM is unknown. For simplicity, presynaptic release machinery, which largely differs among sensory receptor cells and is partially unknown, was omitted from the illustration.

Interestingly, whereas RIM PDZ domains interact with CaV2 (2.1 and 2.2) and RIM C2 domains may bind cytosolic II-III loops in α1 subunits of CaV2.2 and 1.2 (weakly), none of this applies to CaV1.3 channels (72, 172). But, recently, a novel interaction of RIM C2B domain with the COOH terminus of CaV1.3α1 subunit was proposed (278). Thus CaV channels at hair cell ribbon synapses might also bind RIM directly, however, at different interaction sites as their counterparts at conventional synapses. The interactions of RIM with CaV1.4 are not known. Still, the interactions of RIMs and CaV channels also are mediated by other proteins, like β subunits and RIM-BP. RIM C2B domain interacts with β4, β3 and β2a subunits (117, 182). And, the proline-rich sequences (PxxP) of RIM bind to RIM-BPs (395), which in turn interact with CaV1, 2.1 and 2.2 channels (147). RIM proteins also may interact with other synaptic proteins like ELKS/CAST, SNAP-25, synaptotagmin-1, liprin-α, and RAB3. Thus RIMs are well positioned to recruit CaV channels into the AZ (172), although the absence of many of these proteins from mature IHC (but not photoreceptor/bipolar cell) ribbon synapses means that their interactions with RIM may not be relevant for IHC function (e.g., synaptotagmin-1, SNAP-25 etc.; reviewed in Ref. 272).

IHCs lack RIM1, and of the other long RIM isoforms, RIM2α and β are present, with RIM2α being the more abundant one (170). Additionally, IHCs seem to express the short isoform RIM3γ and potentially some RIM2γ, although the function of these isoforms is unclear. Although RIM2α and RIM3γ both promote larger Ca2+ currents in heterologous systems expressing IHC-like Ca2+ channel complexes, the absence of RIM3γ does not significantly affect IHC synaptic function or hearing (278). Consistent with the proposed interactions between long RIM isoforms and the presynaptic AZ cytomatrix, RIM2 isoforms are detected at the base of IHC ribbons, where they are necessary for clustering of CaV1.3 channels (170). In the absence of RIM2α (or both RIM2α and β), CaV channel abundance is reduced throughout the cell but most significantly at the AZs, but CaV channel cluster shape and organization at the AZs remain intact. Accordingly, Ca currents and exocytosis also are reduced by about half but not eliminated (170).

Of note, additional molecular links assist in clustering CaV1.3 near vesicles at IHC AZs (FIGURE 4). These might include palmitoylated β2a subunits that pin trafficked CaV channels to the plasma membrane where RIM-BP might hold them close to the ribbon (also independent from the interaction with RIM). Experiments with RIM-BP2-deficient mice showed that in IHCs RIM-BP2 positively regulates the number of synaptic CaV1.3 channels, and in addition, it supports fast vesicle recruitment into the RRP after depletion, potentially by rapidly bringing synaptic vesicles in close proximity of CaV channels (195). Candidate interaction partners of RIM-BP are the presynaptic scaffold protein bassoon (74; see also below) and ELKS, whose homolog Bruchpilot was detected to interact with RIM-BP at the Drosophila neuromuscular junction (214). The roles of ELKS and CAST, however, are not well understood although they can alter CaV channel properties, as demonstrated at inhibitory hippocampal synapses (213) and in a heterologous expression system (181). Recent data also suggest partially redundant scaffolding roles among ELKS and RIM, at least at central neuronal synapses (391).

At photoreceptor ribbon synapses (394), RIMs have been suggested to serve as AZ scaffolds to which CaV channels are anchored by RIM-BPs (147). Deletion of RIM1/2 reduces presynaptic Ca currents in and, consequently, exocytosis from rod photoreceptors (124). This observation is consistent with the effects of RIM2 deletion observed in IHCs (above), although at retinal ribbon synapses the Ca current reduction seems to be attributable to altered single-channel conductance as CaV channel abundance is not changed (124). Also similar to IHCs, photoreceptor ribbon synapses abundantly express RIM2α, but no RIM1 (124, 218).

In addition to RIM proteins, the large multidomain protein bassoon, together with the ribbon itself, appears to be central to proper CaV channel clustering at the AZs of hair cells and photoreceptors (111, 178, 221, 233, 258, 325). Bassoon binds RIBEYE, a major component of the synaptic ribbon (see Introduction), and anchors the ribbon to the presynaptic density via an unknown molecular link (84; reviewed in Ref. 135). Although bassoon does not interact directly with CaV1.3 (111), disruption of bassoon (and concomitant loss of ribbons from the AZs) results in disorganized CaV channel presynaptic clusters, reduced abundance of CaV channels causing decreased synaptic Ca2+ influx, and fewer docked vesicles. Together, this results in diminished synaptic exocytosis leading to increased SGN spike latencies and desynchronized neural responses to sound onset in the auditory system (48, 111), and abnormal ERG responses in the visual system (84).

Because alterations in bassoon expression result in the loss of ribbons from a majority of AZs, the consequences of bassoon manipulation on CaV channel organization cannot be distinguished from the effects of ribbon depletion. In IHCs, AZs of bassoon mutants that retain ribbons exhibit an intermediate phenotype between wild-type and ribbonless AZs, suggesting that bassoon and ribbon might work in concert (111, 163).

With regard to RIBEYE/Ribeye, evidence from studies of zebrafish, but not mammalian, hair cell ribbon synapses implicates it as required for proper trafficking of CaV channels to the AZs (220, 221, 324, 325). In morpholino zebrafish hair cells, the knock-down of ribeye b (which also significantly reduces levels of Ribeye a) or of both copies of ribeye disrupts the clustering of CaV1.3a channels at the presynaptic AZ and partially affects the tight apposition of postsynaptic afferent contacts, thereby perturbing hearing and vestibular function (325). Mislocalized CaV channels were also found in another study in mutant zebrafish neuromasts with little Ribeye a and no detectable Ribeye b immunofluorescence (221). Conversely, the overexpression of Ribeye leads to formation of ectopic AZs containing ribbons and CaV channel clusters, but lacking postsynaptic partners (325). Enlarged ribbons at innervated AZs, however, do not seem to recruit additional CaV channels, suggesting that other factors may co-determine synaptic CaV channel abundance (323). In immature neuromasts, the effects between CaV channels and ribbon are mutual. Reduction of presynaptic Ca currents (by genetic or pharmacological manipulation) leads first to enlarged and malformed ribbons followed by degradation of synapses (324). Conversely, pharmacologically increased Ca2+ influx decreases the ribbon size and reduces the ribbon number (324).

At mammalian hair cell ribbon synapses, RIBEYE is unlikely to be required for proper trafficking of CaV channels to AZs. First, the distribution of CaV channels in maturing IHCs is broader than that of RIBEYE, suggesting that RIBEYE does not directly recruit CaV channels (402). Furthermore, the abundance of synaptic CaV channels as well as Ca2+ influx are normal in the IHCs of a Ribeye KO mouse model (26, 161), but CaV channel properties become mildly affected in more mature IHCs (161). A depolarized shift of the CaV1.3 activation range in IHCs may partially underlie reduced spontaneous and evoked SGN spiking rates and increased thresholds of sound-evoked spiking in SGNs of the Ribeye KO mice (161). Still, the overall effect of a congenital RIBEYE/ribbon loss has surprisingly mild effects on sound encoding and hearing (26, 161). This may be due to compensatory mechanisms in the IHCs of the Ribeye KO animals, where larger postsynaptic GluA patches (26), and multiple small presynaptic structures resembling conventional AZs were found (161; but see Ref. 26). Likely, a more suitable model to study the role of ribbon in sensory cells would be an inducible KO of RIBEYE. Still, data on constitutive Ribeye deletion helped to partially isolate the roles that bassoon and RIBEYE play individually (26, 161). A difference in the synaptic CaV channel abundance in the IHCs of Ribeye KO versus bassoon mutant mice suggests that bassoon, but not the ribbon, acts to promote channel tethering at the AZ and/or stabilize channels at the AZ (see also Ref. 258); possible mechanisms for channel stabilization include slowing channel recycling and/or degradation and preventing lateral displacement in the face of high rates of membrane turnover that accompany sustained transmission (271). In line with this idea, regulation of the presynaptic ubiquitin-proteasome system (UPS) by bassoon and piccolo, an AZ protein with large similarity to bassoon (103), recently has been demonstrated at CNS synapses (387). In Ribeye KO mouse retinas, Ca2+ influx in bipolar cells seemed unperturbed, but presynaptic CaV channel clusters in photoreceptors (assessed immunohistochemically) were reduced in size (233), contrary to the observation in the hair cells. This suggests RIBEYE may fulfill partially different roles in different cell types and organs or as mentioned above the effects of ribbon loss in IHCs may be partially masked by compensation.

Finally, it also is worth considering possible roles played by other AZ proteins. Ribbon synapses express piccolino, a short isoform of piccolo that lacks several interaction domains (292). Although the absence of piccolino affects ribbon size and shape significantly (291), effects on presynaptic CaV channels remain to be examined. But, its localization towards the top of the ribbon suggests that piccolino might not be an interaction partner of CaV channels (85). Possible roles for ELKS/CAST at ribbon synapses also remain enigmatic. Retinal ribbon synapses are affected by CAST deletion—ribbons are smaller and the CaV channel complement is altered—but effects on CaV channel function have not been ascertained (367c).

3. Positioning CaV channels close to synaptic vesicles

Once CaV channels are targeted to the presynaptic membrane at ribbon synapses, they are held in place very close to release sites to promote nanodomain Ca2+-exocytosis coupling. The ribbon itself could serve a tethering function in the retina (233), but coupling defects are not observed in IHCs of the two bassoon mutants despite defective (or lost) ribbons (111, 163). In IHCs of the RIM-deficient animals, CaV channel coupling to release sites is normal, as suggested by experiments manipulating the number of open channels (170), which is in contrast to the observation made at the calyx of Held (140). Thus RIM-independent molecular links must be of particular importance at ribbon synapses. Possible candidates for bringing CaV1 channels in the proximity of release sites include auxiliary α2δ (153) and β subunits (117, 182) and RIM-BP (4, 74, 128) (FIGURE 4). An additional role in channel positioning may be played by actin filaments, which may hold a subset of vesicles further away from the CaV channels at mature IHC AZs as suggested by recent experiments (134).

Overall, it is easier to exclude than to identify constituents of the protein tethers of CaV1 channels and their molecular links to synaptic vesicles at ribbon synapses. Although studies of the calyx of Held (57) suggest that complexins might influence the distance between CaV channels and synaptic vesicles, this may not be the case at retinal ribbon synapses, where the absence of complexin 3 reduces nanodomain control of exocytosis without affecting Ca current density and presynaptic vesicle distribution (252). It, however, is difficult to determine whether the alteration in nanodomain coupling reflects subtle alterations in the relative placement of channels and vesicles or whether the observation is attributable solely to the effects of complexin 3 on the release machinery itself (252). Furthermore, complexins are not present at IHC synapses (345). Otoferlin, an essential protein for IHC synaptic transmission (307), is also unlikely to influence the CaV channel-synaptic vesicle coupling because otoferlin-deficient IHCs display unperturbed Ca currents (307) and electrophysiology data from otoferlin mutant IHCs suggest unchanged coupling (271, 344). Septin 5 (see sect. IIIB) and synapsin (130) act at conventional synapses to maintain the distance between CaV channels and release sites; however, the latter does not appear to be found in hair cells (235, 377) nor in photoreceptors and bipolar cells (223).

Lastly, it is interesting to consider how synaptic activity itself might alter CaV channel-release site coupling dynamically. As vesicular membrane is added to the plasma membrane during exocytosis and removed during endocytosis, the free intracellular volume of the AZ might change, as might the positions of CaV channels relative to vesicles. Indeed, lateral movement of CaV channels during exocytosis has been observed at photoreceptor AZs (240, 363), and the placement of CaV channels relative to synaptic ribbons can have significant effects on [Ca2+] nanodomains (129). It is interesting to consider the possibility that CaV channel-release site coupling becomes slightly less efficient during periods of sustained Ca2+ entry, but any decrease in coupling efficiency should be countered by accumulation of Ca2+ in the presynaptic terminal.

4. Control of CaV channel degradation

To maintain the functional integrity of the presynaptic AZ, not only must CaV channels be inserted into the membrane correctly, but their removal and degradation must be controlled precisely. CaV channels undergo constant turnover, being internalized and then recycled or degraded to be replaced by newly synthesized ones. The turnover rates for CaV1 channels in the plasma membranes of sensory synapses are not known. In cultured neurons, the metabolic turnover rates for CaV2 channels may be a few days (half-life time of on average 3 days), similar to several other (pre)synaptic proteins (71). Perhaps the most critical process regulating CaV channel turnover is protein ubiquitination, which regulates proteasomal or lysosomal degradation of internalized channels. It has already been noted that the β subunit allows CaV1.2 and CaV2.2 to bypass ER-associated degradation before trafficking to the presynaptic membrane (6, 388), although the mechanism is not clear.

The best model for ubiquitination of ion channels is the epithelial sodium channel ENaC, but also other ion channels, including CaV channels, undergo this process (6, 388). Information regarding regulation of CaV channel membrane density by ubiquitination however is sparse (e.g., Refs. 6, 116, 131, 226, 388). In particular, factors that regulate the rate of ubiquination and modify the process otherwise are not clear. The α2δ subunit for example was proposed to stabilize ion channels in the plasma membrane and prevent their degradation, likely via its MIDAS motif (54). For a detailed discussion on the stabilization, lifetime, and regulation of CaV channel internalization and recycling, the readers are referred to other recent reviews (e.g., Refs. 118, 330).

Here, we will highlight a recently identified regulation of CaV1.3 degradation in hair cells by harmonin (131). Harmonin is a PDZ-domain containing scaffolding protein, required for normal hair cell mechanotransduction (133, 245) as a central organizer of the family of Usher proteins in hair cell stereocilia. It is most well-known for its defects causing the Usher syndrome type 1C, a form of hereditary deaf-blindness, in the case of the type 1C manifested as congenital profound deafness, vestibular dysfunction, and delayed onset retinopathy (retinitis pigmentosa). The mechanism of blindness is not well understood. In the inner ear, depending on the particular mutation, defects in harmonin may affect hair bundle morphogenesis (164), the amplitude of mechanotransduction current and the kinetics of its adaptation (245), or the maturation of mechanotransduction machinery (133). Harmonin might further be involved in modulation of cochlear IHC synaptic function (131, 132). In HEK cells, harmonin tags CaV1.3 channels for ubiquitination and alters their functional levels at the cell surface, which likely restrains their availability also at the IHC ribbon synapse (131). This function of harmonin is perturbed in a deaf-circler mouse mutant model, expressing a mutant protein lacking the central coiled-coil and PST (proline, serine, threonine-rich) domains (164) and unable to interact with CaV1.3 channels (131). As the protein was found at only a fraction of IHC synapses, it was also considered as one of the candidates for establishing heterogeneity of presynaptic Ca2+ signals, perhaps via setting a distinct CaV channel abundance across IHC synapses, but see section IV for further discussion on this topic. The function of this protein at the retinal ribbon synapses, also harboring harmonin (295), is not yet entirely understood (200).

Finally, it is worth considering that ubiquitination of different splice isoforms of the same channel may vary. Splice isoforms of CaV2.2 for example show different susceptibility to UPS-mediated degradation (226), and alternative splicing can contribute also to heart failure by promoting proteasomal degradation of CaV1.2 channels in the heart (155). As several splice isoforms of CaV1.3 in hair cells or CaV1.4 in the retina have been described (see sect. II), it will be interesting to examine these issues further. As well, the role of ubiquitination in the maturation of synapses is worth considering, as the abundance and localization of CaV1.3 changes dramatically upon maturation of IHCs (see sect. V). Heterogeneous CaV channel clusters, which are strictly confined to mature hair cell synapses, may be established via different mechanisms involving processes discussed above (e.g., more precise AZ targeting, differential rates of channel ubiquitination within versus outside the AZs, etc.).

IV. HETEROGENEITY OF CALCIUM INFLUX AT NANODOMAIN-COUPLED ACTIVE ZONES

A. Heterogeneity of Spiral Ganglion Neuron Spiking

The mammalian cochlea encodes sound intensities spanning six orders of magnitude. Individual SGNs, however, encode only a fraction of this range such that different intensity bands are carried by different SGNs; similar parsing of the operating range has been observed in a variety of species (208, 209, 355, 400, 413). In the cat, three functional groups of SGNs were identified: neurons with 1) low and 2) medium spontaneous rates and high thresholds preferentially innervating the neural/modiolar side of IHCs, and 3) high spontaneous rate neurons with low thresholds contacting IHCs at the abneural/pillar side, facing the OHCs (208; FIGURE 5).

FIGURE 5.

FIGURE 5.

Presynaptic IHC heterogeneity supports a wide dynamic range of sound encoding. A: response characteristics of an exemplary high- (HSR; red) and low-spontaneous rate (LSR; blue) SGN fiber. Squares emphasize a difference in dynamic range of the two fibers (i.e., a difference in the stimulus level range between 10 and 90% of the maximal driven spiking rate). SGN fibers segregate into groups that primarily innervate the opposite sides of the IHCs (middle). B and C: heterogeneity of presynaptic Ca2+ signals [voltage sensitivity (B) and abundance of CaV channels (C)] may largely contribute to SGN diversification. B: fractional activation of CaV channels in several active zones (AZs) recorded from a single IHC. Note a hyperpolarized shift in the voltage of half-maximal activation (V0.5) in the abneural (red) CaV channel clusters. C: heterogeneity in the amplitude of IHC presynaptic Ca2+ signals as revealed by confocal Ca2+ imaging. D: hypothetical composition of presynaptic IHC CaV clusters. Middle: fractional activation of hypothetical CaV channel clusters with different proportions of two channel types with distinct activation properties. Top left,: abneural AZs, which are on average smaller and contain fewer CaV channels, contain a higher proportion of channels that activate at more hyperpolarized potentials (magenta). Magenta and green channels may represent distinct splice variants and/or channels with distinct interacting partners that modulate their properties. Bottom left: intermediate scenario with equal proportions of channel types. Right: neural AZs may primarily contain channels with more depolarized activation (green). [Data in B are replotted from Ohn et al. (262). C adapted from Meyer et al. (244), with permission from Springer Nature.]

This diversification of SGNs may originate pre- and/or postsynaptically. One IHC contacts several SGNs (∼5–20 depending on the species and tonotopic region), but a vast majority of SGNs receive one input from a single IHC (109, 208, 210). As an IHC is believed to be isopotential, SGN diversity created by a presynaptic mechanism would require that functionally different AZs exist within one IHC (110, 126, 207, 243, 400, 415). Diversity of AZs could arise from experimentally observed heterogeneity in Ca2+ influx owing to CaV channel properties and/or density, heterogeneity in AZ anatomy and composition, and from variability in the dynamics of exocytosis including the extent of multivesicular release (110, 111, 243, 244, 258, 262, 401, 402).

B. Mechanisms Underlying Heterogeneity of Cochlear Presynaptic Ca2+ Signals

In cats, the SGNs with high spontaneous firing rates and low threshold that contact the pillar side of IHCs are postsynaptic to smaller ribbons than the SGNs with lower spontaneous firing rates that contact the modiolar side of IHCs (243; FIGURE 5). Although the segregation of synapse structure and function in small rodents is not as clear-cut as in the cat auditory system, generally similar observations of AZs come from studies of mouse cochlea (207, 262). Because IHC ribbon size is positively correlated with the size of presynaptic CaV channel cluster and with the magnitude of synaptic Ca2+ signal, it is tempting to consider that presynaptic mechanisms largely underlie SGN diversity (110, 262). Further support for this notion comes from the observation of heterogeneity in the voltage dependence of presynaptic IHC CaV1.3 channel opening (FIGURE 5), with the smaller pillar CaV channel clusters exhibiting slightly lower (∼1.5 mV) activation potentials (262). And, as discussed in section V, the variability of presynaptic Ca2+ signals increases upon development through the emergence of AZs with stronger Ca2+ influx (401, 402), coincident with the appearance of high spontaneous rate auditory fibers in mice (401), which however may mostly contact AZs with weaker maximal Ca2+ influx (see next paragraph).

Thus there seem to be at least two levels of presynaptic CaV regulation of individual AZs, CaV channel density and CaV channel properties. A negative shift in voltage sensitivity could partially account for preferentially lower spiking thresholds of SGNs contacting the pillar side of IHCs. It is puzzling, though, why AZs with greater voltage sensitivity are on average smaller (262). One would expect that under conditions of nanodomain control of exocytosis, smaller AZs with smaller CaV channel clusters would exhibit relatively low glutamate release rates. This would in turn lower the spontaneous rates and increase the threshold of SGN spiking, the opposite of what is observed in SGNs contacting the pillar AZs. And contrary, the large neural modiolar AZs would be expected to drive SGN spiking at lower thresholds, in contrast to the effect expected from their low voltage sensitivity. Potentially the larger ribbons with more CaV channels in the modiolar AZs are opposed to synaptic endings with a different composition of postsynaptic glutamate receptor clusters (perhaps containing many low-affinity receptors), which is worth investigating in the future. This way, the dynamic range of spiking would be broadened, as observed in low spontaneous rate, high-threshold SGN fibers (355). The larger modiolar AZs may thus serve the purpose of supporting a higher dynamic range of transmitter release (262).

What factors could (differentially) set the voltage sensitivity of CaV channels within individual AZs? Possible candidates include different CaV channel splice variants with distinct biophysical properties, preferential abundance of specific auxiliary subunits, and diverse complements of CaV channel modulators. As discussed in section II, IHCs express CaV1.3 channel splice variants with short and long COOH termini, and only the latter contain both interacting regulatory domains that determine the channel’s gating kinetics (332). The function of this COOH-terminal modulator can be interrupted by replacing part of the distal COOH-terminal regulatory domain of the long CaV1.3 channels with a hemagglutinin (HA) tag, making the long HA-tagged variant behave functionally like a short CaV1.3 splice variant in a heterologous expression system (312). However, no shift in CaV channel voltage dependence was observed in IHCs of the CaV1.3DCRDHA/HA KI mice (containing the HA-tag in the long CaV1.3 variant; Refs. 262, 312), suggesting that the absence of a functional CTM in an otherwise ‟longˮ channel can still allow the interaction, e.g., with modulatory proteins and/or that alternative splicing is not involved in establishing CaV heterogeneity.

Alternatively, specific CaV channel interaction partners could be differentially distributed among individual AZs of the IHCs. Detected at only a subset of IHC synapses and found to cause a depolarized shift in CaV channel activation (131, 132, 262), harmonin (see also above) seems like a strong candidate for such a function. But, Ca2+ imaging at individual IHC AZs of wild-type and Ush1c deaf-circler mutant animals suggested that harmonin is not required for heterogeneity of presynaptic Ca2+ signals (262). That study, however, identified a different candidate for diversification of CaV channel voltage activation or abundance: GIPC3, a PDZ-domain-containing cytosolic scaffold protein (reviewed in Ref. 174). Mutation in GIPC3 causes human deafness with audiogenic seizures and progressive hearing loss mainly due to defects in hair bundle structure, mechanotransduction, and hair cell K currents (59, 293). The lack of GIPC3 in a mouse model increases whole cell Ca current and causes an overall shift in voltage activation towards more negative values in IHCs, thereby increasing spontaneous SGN spike rates. More importantly, it disrupts the modiolar-pillar gradient in AZ Ca2+ signal amplitude: in the absence of GIPC3, pillar AZs on average show more Ca2+ influx than AZs facing the modiolus (262). Because the spatial gradient in ribbon/Ca cluster size distributions may be important for generating the broad dynamic range of spiking in low spontaneous rate SGN fibers, it is notable that disruption of GIPC3 narrows the dynamic range of SGN spiking (262). This could be explained by a shift in hearing threshold due to impaired mechanotransduction and cochlear amplification, but the effect also may result, at least in part, from the absence of large AZs at the modiolar side of IHCs. Thus GIPC3 may be one of the factors establishing the heterogeneity of IHC presynaptic AZs, influencing the spatial gradient in CaV channel abundance but not in the Ca current voltage-dependence by an unknown mechanism. By analogy to the related protein GIPC1, which is involved in the establishment of planar cell polarity (PCP) during inner ear epithelial development (119), GIPC3 may assist the spatially polarized trafficking of specific components of the CaV channel complex to or away from the AZs. In the future, it will be interesting to determine whether other proteins relevant for PCP are involved in establishing IHC synaptic heterogeneity.

Another known factor that contributes to creating heterogeneity of maximal Ca2+ influx is bassoon (and/or ribbon itself) as disruption of bassoon results in the absence of large presynaptic Ca clusters (163, 401). As discussed in section IIID2, this may be directly related to the mutual influence between ribbon and CaV channel cluster (324), although this relation may be somewhat different in hair cells of lower or higher vertebrates. Interestingly, the spatial gradient of ribbon sizes also is degraded by noise trauma that causes a temporary hearing threshold shift, and the gradient only partially recovers within a week after the insult (206). The mechanisms behind this phenomenon are not well understood but might include the same processes that establish and maintain the spatial gradients in the healthy cochlea. Most importantly, it remains to be answered how spatial gradient in the voltage dependence of presynaptic Ca2+ influx is established, as this seems to be the most significant determinant of the postsynaptic SGN spiking rates.

C. Role for Heterogeneity at Retinal Synapses

The vertebrate retina faces similar signal processing challenges as it encodes visual stimuli that vary over ∼9 log units in intensity (300). Retinal circuitry has been the subject of many recent reviews to which the reader may refer (e.g., Refs. 14, 77, 231, 263), so discussion here will be constrained narrowly to the issue of variability in [Ca2+] and in Ca2+-exocytosis coupling.

Like those of the cochlea, the retinal signal transducers (rod and cone photoreceptors) distribute their outputs to multiple types of projection neurons, which exhibit varying responses to a sensory stimulus (13). The retina, however, differs from the cochlea in several significant ways that are likely to make synapse-level heterogeneities far less significant for physiological signal processing.

Three features of photoreceptor synapses minimize the effect of quantal variability (both in the amplitude and in the timing of individual exocytotic events) at photoreceptor synapses. 1) Photoreceptors are depolarized in darkness and hyperpolarize in response to light. Thus, in darkness, photoreceptor terminals experience sustained Ca2+ influx that drives tonic exocytosis of glutamatergic vesicles; light is encoded by a decrease rather than an increase in transmitter release (opposite to the coding of sound by IHCs). Consequently, the tonic release rate in darkness is quite high so that occasional slowing of the release rate, as occurs in a stochastic process, is not read out by the downstream circuitry, which has a relatively long integration time, as the absorption of photons (68, 159, 283, 287). Additionally, the long integration times of photoreceptors (10–15 to >100 ms) minimize contributions of heterogeneities in Ca2+ signaling that occur on shorter time scales. 2) Postsynaptic mechanisms act to reduce the influence of individual release events. Postsynaptic signaling in ON bipolar cells is mediated by a signaling cascade that begins with a metabotropic glutamate receptor (mGluR6) and ends with a cation conductance (likely TRPM1), and saturation of the second messenger cascade within this pathway can serve to reduce or eliminate the influence of presynaptic variability on postsynaptic responses (104, 310, 368, 379a). OFF bipolar cells express ionotropic glutamate receptors (both AMPARs and KARs), and the majority of these receptors are likely to be desensitized and/or saturated by tonic glutamate release (78, 80, 125). 3) The morphology of photoreceptor synapses is such that diffusion within and outside of the synaptic cleft (actually, a large synaptic invagination) filters quantal variability (80, 284, 285, 353).

Additionally, retinal circuits are characterized by significant synaptic convergence that reduces the influence that any single synapse could have on downstream neurons. With the exception of the midget pathway in the primate fovea, in which a single cone photoreceptor contacts only one second-order cone bipolar cell that in turn makes synapses with one ganglion cell (187), bipolar cells pool the outputs of multiple photoreceptors and ganglion cells, in turn, receive synapses from many bipolar cells (340). Thus a single ganglion cell can pool the outputs of >10,000 photoreceptors as in the case of convergence of rods to ON alpha ganglion cells in the mammalian retina (76, 340, 373); clearly, heterogeneities in transmission at individual synapses will be averaged out by this circuit design. Nevertheless, Ca2+ signals at individual ribbon synapses in cones have been found to vary strongly, which might help encoding contrasts over a wide range of light intensities (127).

V. DEVELOPMENTAL CHANGES IN CALCIUM CHANNELS AND CALCIUM CHANNEL-EXOCYTOSIS COUPLING

A. CaV Channels in Development of Sensory Receptor Cells and Neural Circuits

CaV channels play a vital role in the development of sensory circuits. Even in the absence of external stimulation, developing sensory systems, like other developing neural circuits, are highly active. Endogenously generated patterned activity might promote neuronal survival, alter or strengthen synaptic connections, and guide the refinement of precise topographic organization of the cortical and subcortical sensory centers or sensory circuits (reviewed in Refs. 3, 38, 53, 173, 201). In the developing cochlea, correlated Ca spikes have been detected in the neighboring IHCs (166, 196, 229, 371, 372). This bursting activity results from interplay between voltage-gated Ca (and Na) and K channels (34, 44, 196, 227229) and may be modulated by ATP released from supporting cells in the cochlear Kölliker’s organ (166, 372, 389; but see Refs. 165, 320). The firing of SGNs and downstream neurons requires synaptic input from pre-hearing cochlear IHCs (121, 370, 371). In the retina, interneurons rather than sensory receptors are responsible for generating spontaneous waves of activity that propagate across the layer of ganglion cells and into subcortical structures (2, 19, 102, 114, 239). However, in both IHCs and photoreceptors, CaV channels are required for proper development and maturation of ribbon synapses (44, 215, 259).

In the absence of CaV1.3 channels, immature IHCs lose the ability to fire Ca spikes and show a strong impairment in development (44). This includes the lack of timely expression of large-conductance Ca2+-activated K (BK) channels and the persistence of immature SK2 channels and direct efferent innervation of IHCs (44, 259). Furthermore, IHCs display several features of morphological immaturity: CaV channel cluster and ribbon size, shape, and distribution appear immature (see below); similarly altered synapses are observed in photoreceptors lacking CaV1.4 (see sect. VIB). In IHCs, this maturation process appears to be controlled by thyroid hormone (TH)-mediated signaling (45, 321).

B. Maturation of IHC Ca2+ Signaling

As IHCs mature, their spontaneous firing activity becomes more structured, exhibiting stereotyped bursts of Ca APs (320), and IHCs change from “pacemakers” that time SGN bursting (371) into sound transducers (34, 120, 196, 401). In mice, this transition occurs around postnatal days 11–12, upon opening of the ear canal and regression of Kölliker’s organ as its spontaneous activity ceases (370). It is believed to be enabled by changes in IHC ionic conductances. BK channels appear, and slow outward-rectifying K channels are upregulated to attenuate spike generation (196, 228). As well, the total number of CaV channels is reduced (34) and their activation kinetics speeds up (417, 418). The remaining CaV channels are confined to the presynaptic AZs, forming clusters (43, 401, 402). Some of those AZs now host larger CaV channel complements, and consequently, stronger local Ca2+ signals emerge (401). As revealed by immunohistochemistry, these clusters undergo intensive remodeling, transforming from spotlike presynaptic arrays to stripes flanking bassoon (402).

The stabilizing interactions with scaffold proteins that support such reorganization are not known, but the proteins bassoon, RIM2, or RIM-BP may be involved, as mouse models with mutations in these proteins show perturbed CaV channel spatial organization and/or abundance (as discussed above; Refs. 111, 170, 195). In the remodeled clusters, CaV channels develop tighter spatial coupling to vesicular sensors of exocytosis (402). Although it remains to be determined how the maturation of CaV channel-vesicle coupling occurs, it clearly has the effect of increasing the Ca2+ efficiency of exocytosis and reducing the cytosolic Ca2+ load accompanying synaptic transmission, thereby lowering the metabolic cost of Ca2+ clearance. Changes in CaV channel clustering are accompanied by concomitant alterations in ribbon size and structure (402). Upon development, the number of ribbons decreases as synapses either undergo fusion or pruning (again, a detailed understanding of underlying processes and mechanisms is lacking). The remaining ribbons are larger and tether more vesicles (402). At the same time, SGNs with higher spontaneous rates and auditory sensitivity can first be detected (401). AZs thus become largely heterogeneous in size and structure (110). This large variability in ribbon sizes and CaV channel clusters likely enables differential coding of auditory information and supports its large dynamic range (described in sect. IV).

VI. CHANNELOPATHIES RELATED TO GENETIC ALTERATIONS OF VOLTAGE-GATED CALCIUM CHANNELS AND INTERACTING PROTEINS

A. Synaptopathies in Hair Cells

To date, two forms of human Ca channelopathies affecting hearing have been identified: one arises from a mutation in the CaV1.3 channel itself (15) and the other from a mutated CaV1.3 modulator, CaBP2 (318; FIGURES 6 and 7). The pathological mutation in CACNA1D, the gene encoding the CaV1.3α1 subunit, was found in a genetic screen of patients with autosomal recessive deafness (15). The affected members of two consanguineous Pakistani families carry a homozygous 3-bp insertion, c.1208_1209insGGG, in the alternatively spliced exon 8B, which is abundantly present in the cochlear IHCs and the pacemaker cells of the sinoatrial node. This in-frame introduction of an additional glycine residue within a highly conserved region near the channel pore results in nonconducting CaV channels (FIGURE 6). The loss of conductivity may be due to abnormal voltage-sensor movements that fail to induce proper pore opening or to occlusion of the channel pore (15). In accordance with the expression pattern of the affected CaV1.3 spliced variant, the major symptoms of CaV1.3 deficiency in human and KO mice are sinoatrial node dysfunction and deafness (15, 280); the channelopathy therefore is termed SANDD syndrome (OMIM 614896). Whether other CACNA1D mutations or polymorphisms contribute to the risk for hearing disorders or heart dysfunction is not known.

FIGURE 6.

FIGURE 6.

Human Ca channelopathies caused by mutations in CaV channels and their interaction partners at ribbon synapses. A: pathogenic loss-of-function mutations causing congenital stationary night blindness type 2 (CSNB2) were found throughout the CaV1.4 structure (red), including the CTM region (orange). At the inner mouth of the channel pore, few mutations were identified to cause a gain-of-function (green circles). B: interestingly, the position of one of the mutations in the CaV1.4 (i.e., in the IS6) coincides with the position of the loss-of-function mutation in the CaV1.3 channel, causing sinoatrial node dysfunction and deafness (SANDD) syndrome. In humans, the function of ribbon presynaptic CaV channels can be further affected by mutations in CaBPs and α2δ4 subunit (red crosses). Note that one intronic CaBP2 base exchange likely leads to alternative splicing, the generation of nonsense codon, and consequently to nonsense-mediated mRNA decay. CaBPs were suggested to prevent Ca2+/CaM binding to the CaV channels and thus reduce calcium-dependent inactivation in both CaV1.4 (A) and CaV1.3 (B) channels. CaBP2 may also affect VDI, potentially shielding the CaV channel pore from the “inactivation lid” (CaVβ-AID complex in the I–II linker).

FIGURE 7.

FIGURE 7.

Model of CaBP2 channelopathy at the inner hair cell ribbon synapse. A: the illustration depicts a ribbon AZ in a CaBP2-deficient compared with a wild-type inner hair cell (IHC) at some time point after the stimulus onset (sine wave). CaBP2 deficiency is proposed to lead to pronounced steady-state CaV channel inactivation. Increased fraction of inactivated channels (magenta) does not support sufficient synaptic transmission: fewer CaV channels are available for triggering release of synaptic vesicles, which is reflected in the spiking pattern of auditory nerve fibers (B). B: representative examples of auditory nerve fiber spike times (wild-type, black-gray; CaBP2-KO, magenta) in response to repetitions of 50-ms tone bursts displayed in dot raster plots. Onset responses are shown with enlarged and darker symbols. Note increased latency and jitter of the first spike as well as decreased evoked spike rates. The examples were taken from data acquired for Picher et al. (277).

Two mutations in the gene coding for CaBP2 cause an autosomal recessive, moderate to severe, prelingual hearing impairment (277, 318) that underlies the deafness locus DFNB93 (354). One mutation, detected in three consanguineous Iranian families, is a splice-site mutation, predicted to cause exon-skipping and produce a truncated protein, lacking the two COOH-terminal EF-hand motifs thus leaving only one functional EF-hand (138). The second mutation, found in an Italian family, leads to nonsense-mediated RNA decay (277). Both result in impaired modulation of CaV1.3 channels, and the hearing defect was recapitulated in a recently generated mouse KO model (277).

CaBP2 might suppress both VDI (preferentially) and CDI (277); the exact mechanism of action however requires further investigation, as does the question of potential partial compensation among different CaBP isoforms present in the IHCs (73). In this context, it is also worth noting that CaBP1, too, has been shown to affect both CDI and VDI in CaV1.2 channels (266, 423) and in heterologously expressed CaV1.3 channels (73). The current hypothesis suggests that the lack (or truncation) of CaBP2 results in a significant steady-state CaV1.3 inactivation, thereby reducing the [Ca2+] necessary to drive efficient synaptic transmission at IHC ribbon synapses (FIGURE 7). CaBP2 might additionally act as a cytosolic buffer, although currently it is not clear if and how a defect in CaBP2-buffering would contribute to the hearing impairment (277). Although the protein is also found in OHCs, its presence does not appear to be required to permit cochlear amplification as evidenced by normal otoacoustic emissions in KO mice and young patients (277; but see Ref. 318).

B. Synaptic Pathologies in Retinal Neurons

Numerous mutations in CaV1.4 (CACNA1F gene; Refs. 25, 349), α2δ4 subunit (CACNA2d-4 gene; Ref. 405; see also sect. IIID1), and CaBP4 (Ref. 420; see also sect. IIB2) are associated with CSNB2 in humans (421; FIGURE 6). These mutations affect retinal proteins in photoreceptors and alter both ON and OFF bipolar signaling pathways. CSNB2 patients show low visual acuity and various degrees of nystagmus, strabismus, and refractive error (37). Electroretinograms as the only precise functional phenotyping and diagnosis criteria of CSNB2 in patients show that both scotopic and photopic responses are affected (421). Their similar phenotypes have recently been reviewed in detail (421), thus we focus here on mutations in the CACNA1F gene as it is most commonly affected in CSNB2.

While most of the few cases associated with CaBP4 mutations present with high hyperopia, refractive errors due to CACNA1F mutations are more variable and include myopia and hyperopia (37, 212). Visual fields in patients are normal, but daylight vision, color vision, and visual acuity can be affected (37). More than 50% of the patients suffer from photophobia (37), often seen in cone dysfunction syndromes (1). Patients with CACNA1F mutations in particular may present with few or no night vision problems (246, 421). Clearly some of the variation in the clinical manifestation of the disease might arise from the different mutations in the CACNA1F gene (FIGURE 6) that cause different channel defects. Judging from studies in various heterologous expression systems, the spectrum of CaV1.4 dysfunction is indeed wide (49, 146, 150, 151, 236, 275, 333). Many mutations cause severe structural changes in the protein so that functional channels are unlikely to form, and most missense mutations lead to dramatically reduced or abolished CaV1.4 activity (for review, see Ref. 342). A typical loss-of-function phenotype might be attributed to fewer functional channels as some of the mutations are predicted to decrease channel stability and promote misfolding (49, 342). Truncation mutations in the COOH terminus were reported to result in the loss of functional CTM and thus unmasked CDI (49, 333). Such mutations may fail to support continuous Ca2+ influx and limit the dynamic range of photoreceptors (e.g., when adapting to variation in illumination).

Our understanding of the pathology of CSNB2 largely stems from various mouse models. The two CaV1.4 KO models available (225, 338) provided the first critical clues; however, they exhibit a much stronger phenotype than human patients (225). CaV1.4 KO mice are functionally blind, show a substantial loss in ganglion cell responsiveness to physiological light stimuli (186), and lack visually evoked cortical activation (225). More similar to the human CSNB2 is a phenotype of a naturally occurring mouse model, nob-2 (56, 91), in which the full-length CaV1.4 protein (carrying a slight modification at the NH2 terminus) is expressed at significantly reduced levels (91). These mice show detectable but reduced ERG responses, although spatial contrast sensitivity measured from optokinetic responses is rather similar to that of wild-type mice (91).

One missense mutation found in a New Zealand family (CaV1.4-I745T in human, mouse homolog I756T; CaV1.4-IT; Ref. 152) has attracted particular interest in the CSNB2 community (FIGURE 8). In a heterologous expression system, CaV1.4-IT gating properties show a remarkable −30 mV shift in the voltage dependence of activation, likely due to increases in CaV1.4 channel conductance and open probability and to significantly slower inactivation kinetics (146). Mice that possess the corresponding mutation recapitulate many aspects of the phenotype of human patients with the same mutation (185, 215, 288) and additionally show a progressive retinal degeneration (186, 288). Judged from ERG recordings, photoreceptor degeneration in the New Zealand family seems rather unlikely (152), but no long-term clinical investigation was undertaken to support this. CSNB2 patients and myopic controls, however, were found by a different study using spectral-domain optical coherence tomography to show differences in retinal layer thickness (63).

FIGURE 8.

FIGURE 8.

Model of CaV1.4 channelopathy at photoreceptor synapse. A: photoreceptor terminals of CaV1.4 wild-type (WT, black) compared with mutant photoreceptors. The gain-of-function mutation CaV1.4-IT (IT, red) is characterized by a strong hyperpolarized shift in the current-voltage relationship (middle). Morphological changes reported in CaV1.4-IT retinas are indicated (right). A certain amount of functional ribbon synapses is still formed in CaV1.4-IT animals; however, these only allow anomalous downstream synaptic transmission (‟delayedˮ). Most synapses in CaV1.4-IT retinas contain immature-like ribbons that may not support synaptic transmission (‟nonresponderˮ). B: representative examples of OFF ganglion cell spike times in response to repetitive light stimuli. C: peri-stimulus time histograms show an increased latency of ganglion cell response of CaV1.4-IT OFF cells.

In the absence of functional CaV1.4 channels, photoreceptor ribbon synapses remain mostly immature, as evidenced by their roundish (or elongated) appearance (215, 286, 288, 414) in the CaV1.4 mutant animal models. Additionally, CaV1.4 dysfunction leads to dendritic sprouting of bipolar and horizontal cells, forming ectopic synapses (185, 215, 288). Similar sprouting of bipolar cell dendrites is seen in mouse models of retinal degeneration in which photoreceptors die (346348). Disruption in the maturation of photoreceptor synaptic ribbons along with free floating ribbons also was observed in cases of changed CaV channel dynamics, e.g., in the presence of CaV1.4-IT (FIGURE 8). Accordingly, the physical integrity of proteins in the ribbon compartment as well as the arciform density/plasma membrane compartment are affected in CaV1.4-IT (288, 338). This phenotype is comparable to the one described for mice lacking intact bassoon, which in photoreceptors and bipolar cells links the two compartments (84, 367a, 290, 337). In addition to its role in anchoring the ribbon, bassoon is also essential in the assembly and transport of ribbon precursor spheres during early steps of photoreceptor synaptogenesis (242, 290, 367b). In some aspects, bassoon-deficient retinas resemble the nob2 (no b-wave 2) phenotype but not that of CaV1.4 KOs. For example, dystrophin, which is a ligand of the dystroglycan/pikachurin complex important for ribbon synapse formation (264), is correctly localized in the nob2 mutants but not in CaV1.4-deficient retinas (24, 338). The small fraction of remaining CaV1.4 channels in the nob2 retinas presumably sustains the binding of some critical interaction partners for the development and (limited) functioning of ribbon synapses. KOs of β2 and α2δ accessory subunits and CaV1.4-activating CaBP4 cause similar defects in synaptic ribbon development and formation (18, 136, 404), supporting the notion that retinal ribbon formation relies on a microdomain environment containing key AZ proteins (289).

C. CaV Channel Pharmacology and Limitations for Targeted Therapies

L-type Ca channel blockers have been widely prescribed for decades as treatment for hypertension and myocardial ischemia. The sensitivity of L-type Ca channels to DHP inhibitors varies between tissues, presumably due to their differential CaV and accessory protein expression. In heterologous expression systems, CaV1.3 and CaV1.4 exhibit ~5- to 10-fold lower sensitivity to DHPs than CaV1.2 at negative membrane potentials (190, 191, 408). This is consistent with the early work on Cav channel pharmacology of photoreceptors, which also suggested a low-affinity blockage of L-type Ca channels (398). The subtype difference might explain why therapeutic doses of DHPs do not cause side effects like changes in hearing thresholds from CaV1.3 inhibition in cochlear IHCs, visual impairment from CaV1.4 block in retinal photoreceptors, and CNS disturbances from CaV1.2/CaV1.3 block in the brain. Moreover, compared with heart tissue, DHP concentrations taken up into the brain were reported to be substantially lower (375). This suggests the uptake through the blood-retina barrier may be restricted. Pharmacological activation of mutated CaV1.4 channels showing a strong reduction in their expression density by L-type Ca channel activators (e.g., BAY K 8644) would not be clinically applicable to human retinal disorders due to toxic side effects expected from activation of CaV1.2 and CaV1.3 in other tissues (for review, see Ref. 419).

VII. CONCLUSIONS

In considering the future of studies of CaV channels and their role in synaptic transmission at ribbon synapses, the establishment and maintenance of AZ organization emerges as an important theme, particularly as “nanodomain” CaV channel-release site coupling has been found to exist at most ribbon synapses. Future work likely will address the following questions: What mechanisms control the trafficking of CaV channels to the AZ, the localization of CaV channels within the AZ, and the lifetimes of the CaV channel proteins?

Although manipulations that dramatically disrupt AZ architecture (e.g., disruption of the organizing protein bassoon and the ribbon protein RIBEYE/Ribeye; Refs. 178, 233, 325) alter surface CaV channel expression, the mechanisms underlying the precise targeting of CaV channels to functional ribbon AZs are unknown. CaV1 lacks a conventional synaptic protein interaction (synprint) site (247), so other, unknown protein-protein interactions must be responsible for the precise CaV channel-release site coupling characteristic of ribbon AZs.

With regard to the localization of CaV channels at the AZ, it has been demonstrated that CaV channels in photoreceptors move—they are displaced laterally—following exocytosis (240). It is of interest, then, to consider how CaV channel-release site coupling can be maintained as large quantities of membrane are added to and removed from the cell surface during tonic exocytosis, particularly as CaV channel mobility at ribbon synapses is affected by membrane composition (241). As well, it is important to recognize that the ability of cells, particularly hair cells, which are tethered in a rigid substrate necessary for mechanotransduction, to deform following exo- and endocytosis is limited.

Finally, although long-term synaptic plasticity at ribbon synapses has not been shown to occur, it likely does given the diurnal variability in presentation of stimuli to sensory organs (e.g., photoreceptor synapse structure changes with the light-dark cycle; Refs. 16, 248). Here, some insight perhaps can be gained from studies of photoreceptors in hibernating mammals or of hair cells exposed to noise trauma (198, 206, 238, 296), where changes in CaV channel abundance or function may be expected.

GRANTS

This work was supported by the German Research Foundation (DFG Priority Program 1608, PA 2769/1–1; to T. Pangrsic), National Eye Institute Grants EY 017836 and 021372 (to J. H. Singer), and Innovative Training Networks 674901 as well as Fonds zur Fӧrderung der wissenschaftlichen Forschung P26881 and P29359 (to A. Koschak).

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

ACKNOWLEDGMENTS

We thank Linda Hsu and Andrej Vilfan for help with the illustrations, Anna Gehrt and Nicola Strenzke for SGN single unit data, Rituparna Chakrabarti for the EM tomography and 3D models of IHC ribbon synapses, Lucia Zanetti for recordings from the retinal ganglion cells, and Regis Nouvian, Jakob Neef, and Maria Magdalena Picher for critical comments and helpful discussions on the manuscript.

Address for reprint requests and other correspondence: T. Pangrsic, Synaptic Physiology of Mammalian Vestibular Hair Cells Group, Institute for Auditory Neuroscience and InnerEarLab, University Medical Center Göttingen, 37099 Göttingen, Germany (e-mail: tpangrs@gwdg.de).

REFERENCES

  • 1.Aboshiha J, Dubis AM, Carroll J, Hardcastle AJ, Michaelides M. The cone dysfunction syndromes. Br J Ophthalmol 100: 115–121, 2016. doi: 10.1136/bjophthalmol-2014-306505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ackman JB, Burbridge TJ, Crair MC. Retinal waves coordinate patterned activity throughout the developing visual system. Nature 490: 219–225, 2012. doi: 10.1038/nature11529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ackman JB, Crair MC. Role of emergent neural activity in visual map development. Curr Opin Neurobiol 24: 166–175, 2014. doi: 10.1016/j.conb.2013.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Acuna C, Liu X, Gonzalez A, Südhof TC. RIM-BPs Mediate Tight Coupling of Action Potentials to Ca(2+)-Triggered Neurotransmitter Release. Neuron 87: 1234–1247, 2015. doi: 10.1016/j.neuron.2015.08.027. [DOI] [PubMed] [Google Scholar]
  • 5.Akopian A, Johnson J, Gabriel R, Brecha N, Witkovsky P. Somatostatin modulates voltage-gated K(+) and Ca(2+) currents in rod and cone photoreceptors of the salamander retina. J Neurosci 20: 929–936, 2000. doi: 10.1523/JNEUROSCI.20-03-00929.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Altier C, Garcia-Caballero A, Simms B, You H, Chen L, Walcher J, Tedford HW, Hermosilla T, Zamponi GW. The Cavβ subunit prevents RFP2-mediated ubiquitination and proteasomal degradation of L-type channels. Nat Neurosci 14: 173–180, 2011. doi: 10.1038/nn.2712. [DOI] [PubMed] [Google Scholar]
  • 7.Anastassov I, Shen W, Ripps H, Chappell RL. Zinc modulation of calcium activity at the photoreceptor terminal: a calcium imaging study. Exp Eye Res 112: 37–44, 2013. doi: 10.1016/j.exer.2013.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Ba-Abbad R, Arno G, Carss K, Stirrups K, Penkett CJ, Moore AT, Michaelides M, Raymond FL, Webster AR, Holder GE. Mutations in CACNA2D4 Cause Distinctive Retinal Dysfunction in Humans. Ophthalmology 123: 668–671.e2, 2016. doi: 10.1016/j.ophtha.2015.09.045. [DOI] [PubMed] [Google Scholar]
  • 9.Babai N, Bartoletti TM, Thoreson WB. Calcium regulates vesicle replenishment at the cone ribbon synapse. J Neurosci 30: 15866–15877, 2010. doi: 10.1523/JNEUROSCI.2891-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Babai N, Sendelbeck A, Regus-Leidig H, Fuchs M, Mertins J, Reim K, Brose N, Feigenspan A, Brandstätter JH. Functional Roles of Complexin 3 and Complexin 4 at Mouse Photoreceptor Ribbon Synapses. J Neurosci 36: 6651–6667, 2016. doi: 10.1523/JNEUROSCI.4335-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Babai N, Thoreson WB. Horizontal cell feedback regulates calcium currents and intracellular calcium levels in rod photoreceptors of salamander and mouse retina. J Physiol 587: 2353–2364, 2009. doi: 10.1113/jphysiol.2009.169656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bacchi N, Messina A, Burtscher V, Dassi E, Provenzano G, Bozzi Y, Demontis GC, Koschak A, Denti MA, Casarosa S. A New Splicing Isoform of Cacna2d4 Mimicking the Effects of c.2451insC Mutation in the Retina: Novel Molecular and Electrophysiological Insights. Invest Ophthalmol Vis Sci 56: 4846–4856, 2015. doi: 10.1167/iovs.15-16410. [DOI] [PubMed] [Google Scholar]
  • 13.Baden T, Berens P, Franke K, Román Rosón M, Bethge M, Euler T. The functional diversity of retinal ganglion cells in the mouse. Nature 529: 345–350, 2016. doi: 10.1038/nature16468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Baden T, Nikolaev A, Esposti F, Dreosti E, Odermatt B, Lagnado L. A synaptic mechanism for temporal filtering of visual signals. PLoS Biol 12: e1001972, 2014. doi: 10.1371/journal.pbio.1001972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Baig SM, Koschak A, Lieb A, Gebhart M, Dafinger C, Nürnberg G, Ali A, Ahmad I, Sinnegger-Brauns MJ, Brandt N, Engel J, Mangoni ME, Farooq M, Khan HU, Nürnberg P, Striessnig J, Bolz HJ. Loss of Ca(v)1.3 (CACNA1D) function in a human channelopathy with bradycardia and congenital deafness. Nat Neurosci 14: 77–84, 2011. doi: 10.1038/nn.2694. [DOI] [PubMed] [Google Scholar]
  • 16.Ball AK, Dickson DH. Diurnal variations in the photoreceptor synaptic terminals of the newt retina. Am J Anat 168: 305–320, 1983. doi: 10.1002/aja.1001680305. [DOI] [PubMed] [Google Scholar]
  • 17.Ball SL, McEnery MW, Yunker AMR, Shin H-S, Gregg RG. Distribution of voltage gated calcium channel β subunits in the mouse retina. Brain Res 1412: 1–8, 2011. doi: 10.1016/j.brainres.2011.07.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Ball SL, Powers PA, Shin HS, Morgans CW, Peachey NS, Gregg RG. Role of the beta(2) subunit of voltage-dependent calcium channels in the retinal outer plexiform layer. Invest Ophthalmol Vis Sci 43: 1595–1603, 2002. [PubMed] [Google Scholar]
  • 19.Bansal A, Singer JH, Hwang BJ, Xu W, Beaudet A, Feller MB. Mice lacking specific nicotinic acetylcholine receptor subunits exhibit dramatically altered spontaneous activity patterns and reveal a limited role for retinal waves in forming ON and OFF circuits in the inner retina. J Neurosci 20: 7672–7681, 2000. doi: 10.1523/JNEUROSCI.20-20-07672.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Bao H, Wong WH, Goldberg JM, Eatock RA. Voltage-gated calcium channel currents in type I and type II hair cells isolated from the rat crista. J Neurophysiol 90: 155–164, 2003. doi: 10.1152/jn.00244.2003. [DOI] [PubMed] [Google Scholar]
  • 21.Barnes S, Merchant V, Mahmud F. Modulation of transmission gain by protons at the photoreceptor output synapse. Proc Natl Acad Sci USA 90: 10081–10085, 1993. doi: 10.1073/pnas.90.21.10081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Bartoletti TM, Jackman SL, Babai N, Mercer AJ, Kramer RH, Thoreson WB. Release from the cone ribbon synapse under bright light conditions can be controlled by the opening of only a few Ca(2+) channels. J Neurophysiol 106: 2922–2935, 2011. doi: 10.1152/jn.00634.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Baumann L, Gerstner A, Zong X, Biel M, Wahl-Schott C. Functional characterization of the L-type Ca2+ channel Cav1.4alpha1 from mouse retina. Invest Ophthalmol Vis Sci 45: 708–713, 2004. doi: 10.1167/iovs.03-0937. [DOI] [PubMed] [Google Scholar]
  • 24.Bayley PR, Morgans CW. Rod bipolar cells and horizontal cells form displaced synaptic contacts with rods in the outer nuclear layer of the nob2 retina. J Comp Neurol 500: 286–298, 2007. doi: 10.1002/cne.21188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Bech-Hansen NT, Naylor MJ, Maybaum TA, Pearce WG, Koop B, Fishman GA, Mets M, Musarella MA, Boycott KM. Loss-of-function mutations in a calcium-channel alpha1-subunit gene in Xp11.23 cause incomplete X-linked congenital stationary night blindness. Nat Genet 19: 264–267, 1998. doi: 10.1038/947. [DOI] [PubMed] [Google Scholar]
  • 26.Becker L, Schnee ME, Niwa M, Sun W, Maxeiner S, Talaei S, Kachar B, Rutherford MA, Ricci AJ. The presynaptic ribbon facilitates sustained exocytosis at the hair cell afferent fiber synape. eLife 7: e30241, 2018. doi: 10.7554/eLife.30241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Beites CL, Xie H, Bowser R, Trimble WS. The septin CDCrel-1 binds syntaxin and inhibits exocytosis. Nat Neurosci 2: 434–439, 1999. doi: 10.1038/8100. [DOI] [PubMed] [Google Scholar]
  • 28.Bemben MA, Shipman SL, Nicoll RA, Roche KW. The cellular and molecular landscape of neuroligins. Trends Neurosci 38: 496–505, 2015. doi: 10.1016/j.tins.2015.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ben Johny M, Yang PS, Bazzazi H, Yue DT. Dynamic switching of calmodulin interactions underlies Ca2+ regulation of CaV1.3 channels. Nat Commun 4: 1717, 2013. doi: 10.1038/ncomms2727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Ben-Johny M, Dick IE, Sang L, Limpitikul WB, Kang PW, Niu J, Banerjee R, Yang W, Babich JS, Issa JB, Lee SR, Namkung H, Li J, Zhang M, Yang PS, Bazzazi H, Adams PJ, Joshi-Mukherjee R, Yue DN, Yue DT. Towards a Unified Theory of Calmodulin Regulation (Calmodulation) of Voltage-Gated Calcium and Sodium Channels. Curr Mol Pharmacol 8: 188–205, 2015. doi: 10.2174/1874467208666150507110359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Ben-Johny M, Yue DT. Calmodulin regulation (calmodulation) of voltage-gated calcium channels. J Gen Physiol 143: 679–692, 2014. doi: 10.1085/jgp.201311153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Bernstein GM, Jones OT. Kinetics of internalization and degradation of N-type voltage-gated calcium channels: role of the alpha2/delta subunit. Cell Calcium 41: 27–40, 2007. doi: 10.1016/j.ceca.2006.04.010. [DOI] [PubMed] [Google Scholar]
  • 33.Berntson A, Taylor WR, Morgans CW. Molecular identity, synaptic localization, and physiology of calcium channels in retinal bipolar cells. J Neurosci Res 71: 146–151, 2003. doi: 10.1002/jnr.10459. [DOI] [PubMed] [Google Scholar]
  • 34.Beutner D, Moser T. The presynaptic function of mouse cochlear inner hair cells during development of hearing. J Neurosci 21: 4593–4599, 2001. doi: 10.1523/JNEUROSCI.21-13-04593.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Beutner D, Voets T, Neher E, Moser T. Calcium dependence of exocytosis and endocytosis at the cochlear inner hair cell afferent synapse. Neuron 29: 681–690, 2001. doi: 10.1016/S0896-6273(01)00243-4. [DOI] [PubMed] [Google Scholar]
  • 36.Bichet D, Cornet V, Geib S, Carlier E, Volsen S, Hoshi T, Mori Y, De Waard M. The I–II loop of the Ca2+ channel alpha1 subunit contains an endoplasmic reticulum retention signal antagonized by the beta subunit. Neuron 25: 177–190, 2000. doi: 10.1016/S0896-6273(00)80881-8. [DOI] [PubMed] [Google Scholar]
  • 37.Bijveld MMC, Florijn RJ, Bergen AAB, van den Born LI, Kamermans M, Prick L, Riemslag FCC, van Schooneveld MJ, Kappers AML, van Genderen MM. Genotype and phenotype of 101 dutch patients with congenital stationary night blindness. Ophthalmology 120: 2072–2081, 2013. doi: 10.1016/j.ophtha.2013.03.002. [DOI] [PubMed] [Google Scholar]
  • 38.Blankenship AG, Feller MB. Mechanisms underlying spontaneous patterned activity in developing neural circuits. Nat Rev Neurosci 11: 18–29, 2010. doi: 10.1038/nrn2759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Bock G, Gebhart M, Scharinger A, Jangsangthong W, Busquet P, Poggiani C, Sartori S, Mangoni ME, Sinnegger-Brauns MJ, Herzig S, Striessnig J, Koschak A. Functional properties of a newly identified C-terminal splice variant of Cav1.3 L-type Ca2+ channels. J Biol Chem 286: 42736–42748, 2011. doi: 10.1074/jbc.M111.269951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Bollmann JH, Sakmann B. Control of synaptic strength and timing by the release-site Ca2+ signal. Nat Neurosci 8: 426–434, 2005. doi: 10.1038/nn1417. [DOI] [PubMed] [Google Scholar]
  • 41.Bollmann JH, Sakmann B, Borst JGG. Calcium sensitivity of glutamate release in a calyx-type terminal. Science 289: 953–957, 2000. doi: 10.1126/science.289.5481.953. [DOI] [PubMed] [Google Scholar]
  • 42.Borst JGG, Sakmann B. Calcium influx and transmitter release in a fast CNS synapse. Nature 383: 431–434, 1996. doi: 10.1038/383431a0. [DOI] [PubMed] [Google Scholar]
  • 43.Brandt A, Khimich D, Moser T. Few CaV1.3 channels regulate the exocytosis of a synaptic vesicle at the hair cell ribbon synapse. J Neurosci 25: 11577–11585, 2005. doi: 10.1523/JNEUROSCI.3411-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Brandt A, Striessnig J, Moser T. CaV1.3 channels are essential for development and presynaptic activity of cochlear inner hair cells. J Neurosci 23: 10832–10840, 2003. doi: 10.1523/JNEUROSCI.23-34-10832.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Brandt N, Kuhn S, Münkner S, Braig C, Winter H, Blin N, Vonthein R, Knipper M, Engel J. Thyroid hormone deficiency affects postnatal spiking activity and expression of Ca2+ and K+ channels in rodent inner hair cells. J Neurosci 27: 3174–3186, 2007. doi: 10.1523/JNEUROSCI.3965-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Briggman KL, Helmstaedter M, Denk W. Wiring specificity in the direction-selectivity circuit of the retina. Nature 471: 183–188, 2011. doi: 10.1038/nature09818. [DOI] [PubMed] [Google Scholar]
  • 47.Buraei Z, Yang J. The β subunit of voltage-gated Ca2+ channels. Physiol Rev 90: 1461–1506, 2010. doi: 10.1152/physrev.00057.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Buran BN, Strenzke N, Neef A, Gundelfinger ED, Moser T, Liberman MC. Onset coding is degraded in auditory nerve fibers from mutant mice lacking synaptic ribbons. J Neurosci 30: 7587–7597, 2010. doi: 10.1523/JNEUROSCI.0389-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Burtscher V, Schicker K, Novikova E, Pöhn B, Stockner T, Kugler C, Singh A, Zeitz C, Lancelot M-E, Audo I, Leroy BP, Freissmuth M, Herzig S, Matthes J, Koschak A. Spectrum of Cav1.4 dysfunction in congenital stationary night blindness type 2. Biochim Biophys Acta 1838: 2053–2065, 2014. doi: 10.1016/j.bbamem.2014.04.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Busquet P, Nguyen NK, Schmid E, Tanimoto N, Seeliger MW, Ben-Yosef T, Mizuno F, Akopian A, Striessnig J, Singewald N. CaV1.3 L-type Ca2+ channels modulate depression-like behaviour in mice independent of deaf phenotype. Int J Neuropsychopharmacol 13: 499–513, 2010. doi: 10.1017/S1461145709990368. [DOI] [PubMed] [Google Scholar]
  • 51.Cadetti L, Bryson EJ, Ciccone CA, Rabl K, Thoreson WB. Calcium-induced calcium release in rod photoreceptor terminals boosts synaptic transmission during maintained depolarization. Eur J Neurosci 23: 2983–2990, 2006. doi: 10.1111/j.1460-9568.2006.04845.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Calin-Jageman I, Lee A. Ca(v)1 L-type Ca2+ channel signaling complexes in neurons. J Neurochem 105: 573–583, 2008. doi: 10.1111/j.1471-4159.2008.05286.x. [DOI] [PubMed] [Google Scholar]
  • 53.Cang J, Feldheim DA. Developmental mechanisms of topographic map formation and alignment. Annu Rev Neurosci 36: 51–77, 2013. doi: 10.1146/annurev-neuro-062012-170341. [DOI] [PubMed] [Google Scholar]
  • 54.Cantí C, Nieto-Rostro M, Foucault I, Heblich F, Wratten J, Richards MW, Hendrich J, Douglas L, Page KM, Davies A, Dolphin AC. The metal-ion-dependent adhesion site in the Von Willebrand factor-A domain of alpha2delta subunits is key to trafficking voltage-gated Ca2+ channels. Proc Natl Acad Sci USA 102: 11230–11235, 2005. doi: 10.1073/pnas.0504183102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol 3: a003947, 2011. doi: 10.1101/cshperspect.a003947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Chang B, Heckenlively JR, Bayley PR, Brecha NC, Davisson MT, Hawes NL, Hirano AA, Hurd RE, Ikeda A, Johnson BA, McCall MA, Morgans CW, Nusinowitz S, Peachey NS, Rice DS, Vessey KA, Gregg RG. The nob2 mouse, a null mutation in Cacna1f: anatomical and functional abnormalities in the outer retina and their consequences on ganglion cell visual responses. Vis Neurosci 23: 11–24, 2006. doi: 10.1017/S095252380623102X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Chang S, Reim K, Pedersen M, Neher E, Brose N, Taschenberger H. Complexin stabilizes newly primed synaptic vesicles and prevents their premature fusion at the mouse calyx of held synapse. J Neurosci 35: 8272–8290, 2015. doi: 10.1523/JNEUROSCI.4841-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Chapochnikov NM, Takago H, Huang C-H, Pangršič T, Khimich D, Neef J, Auge E, Göttfert F, Hell SW, Wichmann C, Wolf F, Moser T. Uniquantal release through a dynamic fusion pore is a candidate mechanism of hair cell exocytosis. Neuron 83: 1389–1403, 2014. doi: 10.1016/j.neuron.2014.08.003. [DOI] [PubMed] [Google Scholar]
  • 59.Charizopoulou N, Lelli A, Schraders M, Ray K, Hildebrand MS, Ramesh A, Srisailapathy CRS, Oostrik J, Admiraal RJC, Neely HR, Latoche JR, Smith RJH, Northup JK, Kremer H, Holt JR, Noben-Trauth K. Gipc3 mutations associated with audiogenic seizures and sensorineural hearing loss in mouse and human. Nat Commun 2: 201, 2011. doi: 10.1038/ncomms1200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Chen M, Van Hook MJ, Thoreson WB. Ca2+ Diffusion through Endoplasmic Reticulum Supports Elevated Intraterminal Ca2+ Levels Needed to Sustain Synaptic Release from Rods in Darkness. J Neurosci 35: 11364–11373, 2015. doi: 10.1523/JNEUROSCI.0754-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Chen M, Van Hook MJ, Zenisek D, Thoreson WB. Properties of ribbon and non-ribbon release from rod photoreceptors revealed by visualizing individual synaptic vesicles. J Neurosci 33: 2071–2086, 2013. doi: 10.1523/JNEUROSCI.3426-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chen M, Križaj D, Thoreson WB. Intracellular calcium stores drive slow non-ribbon vesicle release from rod photoreceptors. Front Cell Neurosci 8: 20, 2014. doi: 10.3389/fncel.2014.00020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Chen RWS, Greenberg JP, Lazow MA, Ramachandran R, Lima LH, Hwang JC, Schubert C, Braunstein A, Allikmets R, Tsang SH. Autofluorescence imaging and spectral-domain optical coherence tomography in incomplete congenital stationary night blindness and comparison with retinitis pigmentosa. Am J Ophthalmol 153: 143–154.e2, 2012. doi: 10.1016/j.ajo.2011.06.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Chen XH, Bezprozvanny I, Tsien RW. Molecular basis of proton block of L-type Ca2+ channels. J Gen Physiol 108: 363–374, 1996. doi: 10.1085/jgp.108.5.363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Chien AJ, Gao T, Perez-Reyes E, Hosey MM. Membrane targeting of L-type calcium channels. Role of palmitoylation in the subcellular localization of the beta2a subunit. J Biol Chem 273: 23590–23597, 1998. doi: 10.1074/jbc.273.36.23590. [DOI] [PubMed] [Google Scholar]
  • 66.Cho S, von Gersdorff H. Proton-mediated block of Ca2+ channels during multivesicular release regulates short-term plasticity at an auditory hair cell synapse. J Neurosci 34: 15877–15887, 2014. doi: 10.1523/JNEUROSCI.2304-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Cho S, Li G-L, von Gersdorff H. Recovery from short-term depression and facilitation is ultrafast and Ca2+ dependent at auditory hair cell synapses. J Neurosci 31: 5682–5692, 2011. doi: 10.1523/JNEUROSCI.5453-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Choi S-Y, Borghuis BG, Rea R, Levitan ES, Sterling P, Kramer RH. Encoding light intensity by the cone photoreceptor synapse. Neuron 48: 555–562, 2005. doi: 10.1016/j.neuron.2005.09.011. [DOI] [PubMed] [Google Scholar]
  • 69.Christel CJ, Schaer R, Wang S, Henzi T, Kreiner L, Grabs D, Schwaller B, Lee A. Calretinin regulates Ca2+-dependent inactivation and facilitation of Ca(v)2.1 Ca2+ channels through a direct interaction with the α12.1 subunit. J Biol Chem 287: 39766–39775, 2012. doi: 10.1074/jbc.M112.406363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Coggins M, Zenisek D. Evidence that exocytosis is driven by calcium entry through multiple calcium channels in goldfish retinal bipolar cells. J Neurophysiol 101: 2601–2619, 2009. doi: 10.1152/jn.90881.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Cohen LD, Zuchman R, Sorokina O, Müller A, Dieterich DC, Armstrong JD, Ziv T, Ziv NE. Metabolic turnover of synaptic proteins: kinetics, interdependencies and implications for synaptic maintenance. PLoS One 8: e63191, 2013. doi: 10.1371/journal.pone.0063191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Coppola T, Magnin-Lüthi S, Perret-Menoud V, Gattesco S, Schiavo G, Regazzi R. Direct interaction of the Rab3 effector RIM with Ca2+ channels, SNAP-25, and synaptotagmin. J Biol Chem 276: 32756–32762, 2001. doi: 10.1074/jbc.M100929200. [DOI] [PubMed] [Google Scholar]
  • 73.Cui G, Meyer AC, Calin-Jageman I, Neef J, Haeseleer F, Moser T, Lee A. Ca2+-binding proteins tune Ca2+-feedback to Cav1.3 channels in mouse auditory hair cells. J Physiol 585: 791–803, 2007. doi: 10.1113/jphysiol.2007.142307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Davydova D, Marini C, King C, Klueva J, Bischof F, Romorini S, Montenegro-Venegas C, Heine M, Schneider R, Schröder MS, Altrock WD, Henneberger C, Rusakov DA, Gundelfinger ED, Fejtova A. Bassoon specifically controls presynaptic P/Q-type Ca(2+) channels via RIM-binding protein. Neuron 82: 181–194, 2014. doi: 10.1016/j.neuron.2014.02.012. [DOI] [PubMed] [Google Scholar]
  • 75.De Sevilla Müller LP, Liu J, Solomon A, Rodriguez A, Brecha NC. Expression of voltage-gated calcium channel α(2)δ(4) subunits in the mouse and rat retina. J Comp Neurol 521: 2486–2501, 2013. doi: 10.1002/cne.23294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Demb JB, Singer JH. Intrinsic properties and functional circuitry of the AII amacrine cell. Vis Neurosci 29: 51–60, 2012. doi: 10.1017/S0952523811000368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Demb JB, Singer JH. Functional Circuitry of the Retina. Annu Rev Vis Sci 1: 263–289, 2015. doi: 10.1146/annurev-vision-082114-035334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.DeVries SH. Bipolar cells use kainate and AMPA receptors to filter visual information into separate channels. Neuron 28: 847–856, 2000. doi: 10.1016/S0896-6273(00)00158-6. [DOI] [PubMed] [Google Scholar]
  • 79.DeVries SH. Exocytosed protons feedback to suppress the Ca2+ current in mammalian cone photoreceptors. Neuron 32: 1107–1117, 2001. doi: 10.1016/S0896-6273(01)00535-9. [DOI] [PubMed] [Google Scholar]
  • 80.DeVries SH, Li W, Saszik S. Parallel processing in two transmitter microenvironments at the cone photoreceptor synapse. Neuron 50: 735–748, 2006. doi: 10.1016/j.neuron.2006.04.034. [DOI] [PubMed] [Google Scholar]
  • 81.Diamond JS, Jahr CE. Asynchronous release of synaptic vesicles determines the time course of the AMPA receptor-mediated EPSC. Neuron 15: 1097–1107, 1995. doi: 10.1016/0896-6273(95)90098-5. [DOI] [PubMed] [Google Scholar]
  • 82.Dick IE, Limpitikul WB, Niu J, Banerjee R, Issa JB, Ben-Johny M, Adams PJ, Kang PW, Lee SR, Sang L, Yang W, Babich J, Zhang M, Bazazzi H, Yue NC, Tomaselli GF. A rendezvous with the queen of ion channels: three decades of ion channel research by David T Yue and his Calcium Signals Laboratory. Channels (Austin) 10: 20–32, 2016. doi: 10.1080/19336950.2015.1051272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Dick IE, Tadross MR, Liang H, Tay LH, Yang W, Yue DT. A modular switch for spatial Ca2+ selectivity in the calmodulin regulation of CaV channels. Nature 451: 830–834, 2008. doi: 10.1038/nature06529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Dick O, tom Dieck S, Altrock WD, Ammermüller J, Weiler R, Garner CC, Gundelfinger ED, Brandstätter JH. The presynaptic active zone protein bassoon is essential for photoreceptor ribbon synapse formation in the retina. Neuron 37: 775–786, 2003. doi: 10.1016/S0896-6273(03)00086-2. [DOI] [PubMed] [Google Scholar]
  • 85.Dick O, Hack I, Altrock WD, Garner CC, Gundelfinger ED, Brandstätter JH. Localization of the presynaptic cytomatrix protein Piccolo at ribbon and conventional synapses in the rat retina: comparison with Bassoon. J Comp Neurol 439: 224–234, 2001. doi: 10.1002/cne.1344. [DOI] [PubMed] [Google Scholar]
  • 87.Dodge FA Jr, Rahamimoff R. Co-operative action a calcium ions in transmitter release at the neuromuscular junction. J Physiol 193: 419–432, 1967. doi: 10.1113/jphysiol.1967.sp008367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Doering CJ, Hamid J, Simms B, McRory JE, Zamponi GW. Cav1.4 encodes a calcium channel with low open probability and unitary conductance. Biophys J 89: 3042–3048, 2005. doi: 10.1529/biophysj.105.067124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Doering CJ, Rehak R, Bonfield S, Peloquin JB, Stell WK, Mema SC, Sauvé Y, McRory JE. Modified Ca(v)1.4 expression in the Cacna1f(nob2) mouse due to alternative splicing of an ETn inserted in exon 2. PLoS One 3: e2538, 2008. doi: 10.1371/journal.pone.0002538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Dolmetsch RE, Pajvani U, Fife K, Spotts JM, Greenberg ME. Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science 294: 333–339, 2001. doi: 10.1126/science.1063395. [DOI] [PubMed] [Google Scholar]
  • 93.Dolphin AC. Calcium channel diversity: multiple roles of calcium channel subunits. Curr Opin Neurobiol 19: 237–244, 2009. doi: 10.1016/j.conb.2009.06.006. [DOI] [PubMed] [Google Scholar]
  • 94.Dolphin AC. Calcium channel auxiliary α2δ and β subunits: trafficking and one step beyond. Nat Rev Neurosci 13: 542–555, 2012. doi: 10.1038/nrn3311. [DOI] [PubMed] [Google Scholar]
  • 95.Dou H, Vazquez AE, Namkung Y, Chu H, Cardell EL, Nie L, Parson S, Shin H-S, Yamoah EN. Null mutation of alpha1D Ca2+ channel gene results in deafness but no vestibular defect in mice. J Assoc Res Otolaryngol 5: 215–226, 2004. doi: 10.1007/s10162-003-4020-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Dulon D, Safieddine S, Jones SM, Petit C. Otoferlin is critical for a highly sensitive and linear calcium-dependent exocytosis at vestibular hair cell ribbon synapses. J Neurosci 29: 10474–10487, 2009. doi: 10.1523/JNEUROSCI.1009-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Duncan G, Rabl K, Gemp I, Heidelberger R, Thoreson WB. Quantitative analysis of synaptic release at the photoreceptor synapse. Biophys J 98: 2102–2110, 2010. doi: 10.1016/j.bpj.2010.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Eggermann E, Bucurenciu I, Goswami SP, Jonas P. Nanodomain coupling between Ca2+ channels and sensors of exocytosis at fast mammalian synapses. Nat Rev Neurosci 13: 7–21, 2012. doi: 10.1038/nrn3125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Erickson MG, Liang H, Mori MX, Yue DT. FRET two-hybrid mapping reveals function and location of L-type Ca2+ channel CaM preassociation. Neuron 39: 97–107, 2003. doi: 10.1016/S0896-6273(03)00395-7. [DOI] [PubMed] [Google Scholar]
  • 100.Faas GC, Raghavachari S, Lisman JE, Mody I. Calmodulin as a direct detector of Ca2+ signals. Nat Neurosci 14: 301–304, 2011. doi: 10.1038/nn.2746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Fell B, Eckrich S, Blum K, Eckrich T, Hecker D, Obermair GJ, Münkner S, Flockerzi V, Schick B, Engel J. α2δ2 Controls the Function and Trans-Synaptic Coupling of Cav1.3 Channels in Mouse Inner Hair Cells and Is Essential for Normal Hearing. J Neurosci 36: 11024–11036, 2016. doi: 10.1523/JNEUROSCI.3468-14.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Feller MB, Wellis DP, Stellwagen D, Werblin FS, Shatz CJ. Requirement for cholinergic synaptic transmission in the propagation of spontaneous retinal waves. Science 272: 1182–1187, 1996. doi: 10.1126/science.272.5265.1182. [DOI] [PubMed] [Google Scholar]
  • 103.Fenster SD, Chung WJ, Zhai R, Cases-Langhoff C, Voss B, Garner AM, Kaempf U, Kindler S, Gundelfinger ED, Garner CC. Piccolo, a presynaptic zinc finger protein structurally related to bassoon. Neuron 25: 203–214, 2000. doi: 10.1016/S0896-6273(00)80883-1. [DOI] [PubMed] [Google Scholar]
  • 104.Field GD, Rieke F. Mechanisms regulating variability of the single photon responses of mammalian rod photoreceptors. Neuron 35: 733–747, 2002. doi: 10.1016/S0896-6273(02)00822-X. [DOI] [PubMed] [Google Scholar]
  • 105.Findeisen F, Minor DL Jr. Progress in the structural understanding of voltage-gated calcium channel (CaV) function and modulation. Channels (Austin) 4: 459–474, 2010. doi: 10.4161/chan.4.6.12867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Findeisen F, Minor DL Jr. Structural basis for the differential effects of CaBP1 and calmodulin on Ca(V)1.2 calcium-dependent inactivation. Structure 18: 1617–1631, 2010. doi: 10.1016/j.str.2010.09.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Findeisen F, Rumpf CH, Minor DL Jr. Apo states of calmodulin and CaBP1 control CaV1 voltage-gated calcium channel function through direct competition for the IQ domain. J Mol Biol 425: 3217–3234, 2013. doi: 10.1016/j.jmb.2013.06.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Findeisen F, Tolia A, Arant R, Kim EY, Isacoff E, Minor DL Jr. Calmodulin overexpression does not alter Cav1.2 function or oligomerization state. Channels (Austin) 5: 320–324, 2011. doi: 10.4161/chan.5.4.16821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Francis HW, Rivas A, Lehar M, Ryugo DK. Two types of afferent terminals innervate cochlear inner hair cells in C57BL/6J mice. Brain Res 1016: 182–194, 2004. doi: 10.1016/j.brainres.2004.05.016. [DOI] [PubMed] [Google Scholar]
  • 110.Frank T, Khimich D, Neef A, Moser T. Mechanisms contributing to synaptic Ca2+ signals and their heterogeneity in hair cells. Proc Natl Acad Sci USA 106: 4483–4488, 2009. doi: 10.1073/pnas.0813213106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Frank T, Rutherford MA, Strenzke N, Neef A, Pangršič T, Khimich D, Fejtova A, Gundelfinger ED, Liberman MC, Harke B, Bryan KE, Lee A, Egner A, Riedel D, Moser T. Bassoon and the synaptic ribbon organize Ca2+ channels and vesicles to add release sites and promote refilling. Neuron 68: 724–738, 2010. doi: 10.1016/j.neuron.2010.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Fuchs PA, Evans MG, Murrow BW. Calcium currents in hair cells isolated from the cochlea of the chick. J Physiol 429: 553–568, 1990. doi: 10.1113/jphysiol.1990.sp018272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Furukawa T, Kuno M, Matsuura S. Quantal analysis of a decremental response at hair cell-afferent fibre synapses in the goldfish sacculus. J Physiol 322: 181–195, 1982. doi: 10.1113/jphysiol.1982.sp014031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Galli L, Maffei L. Spontaneous impulse activity of rat retinal ganglion cells in prenatal life. Science 242: 90–91, 1988. doi: 10.1126/science.3175637. [DOI] [PubMed] [Google Scholar]
  • 115.Gao B, Sekido Y, Maximov A, Saad M, Forgacs E, Latif F, Wei MH, Lerman M, Lee JH, Perez-Reyes E, Bezprozvanny I, Minna JD. Functional properties of a new voltage-dependent calcium channel alpha(2)delta auxiliary subunit gene (CACNA2D2). J Biol Chem 275: 12237–12242, 2000. doi: 10.1074/jbc.275.16.12237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.García-Caballero A, Gadotti VM, Stemkowski P, Weiss N, Souza IA, Hodgkinson V, Bladen C, Chen L, Hamid J, Pizzoccaro A, Deage M, François A, Bourinet E, Zamponi GW. The deubiquitinating enzyme USP5 modulates neuropathic and inflammatory pain by enhancing Cav3.2 channel activity. Neuron 83: 1144–1158, 2014. doi: 10.1016/j.neuron.2014.07.036. [DOI] [PubMed] [Google Scholar]
  • 117.Gebhart M, Juhasz-Vedres G, Zuccotti A, Brandt N, Engel J, Trockenbacher A, Kaur G, Obermair GJ, Knipper M, Koschak A, Striessnig J. Modulation of Cav1.3 Ca2+ channel gating by Rab3 interacting molecule. Mol Cell Neurosci 44: 246–259, 2010. doi: 10.1016/j.mcn.2010.03.011. [DOI] [PubMed] [Google Scholar]
  • 118.Gerrow K, Triller A. Synaptic stability and plasticity in a floating world. Curr Opin Neurobiol 20: 631–639, 2010. doi: 10.1016/j.conb.2010.06.010. [DOI] [PubMed] [Google Scholar]
  • 119.Giese AP, Ezan J, Wang L, Lasvaux L, Lembo F, Mazzocco C, Richard E, Reboul J, Borg J-P, Kelley MW, Sans N, Brigande J, Montcouquiol M. Gipc1 has a dual role in Vangl2 trafficking and hair bundle integrity in the inner ear. Development 139: 3775–3785, 2012. doi: 10.1242/dev.074229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Glowatzki E, Fuchs PA. Cholinergic synaptic inhibition of inner hair cells in the neonatal mammalian cochlea. Science 288: 2366–2368, 2000. doi: 10.1126/science.288.5475.2366. [DOI] [PubMed] [Google Scholar]
  • 121.Glowatzki E, Fuchs PA. Transmitter release at the hair cell ribbon synapse. Nat Neurosci 5: 147–154, 2002. doi: 10.1038/nn796. [DOI] [PubMed] [Google Scholar]
  • 122.Gomez-Ospina N, Tsuruta F, Barreto-Chang O, Hu L, Dolmetsch R. The C terminus of the L-type voltage-gated calcium channel Ca(V)1.2 encodes a transcription factor. Cell 127: 591–606, 2006. doi: 10.1016/j.cell.2006.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Goutman JD, Glowatzki E. Time course and calcium dependence of transmitter release at a single ribbon synapse. Proc Natl Acad Sci USA 104: 16341–16346, 2007. doi: 10.1073/pnas.0705756104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Grabner CP, Gandini MA, Rehak R, Le Y, Zamponi GW, Schmitz F. RIM1/2-Mediated Facilitation of Cav1.4 Channel Opening Is Required for Ca2+-Stimulated Release in Mouse Rod Photoreceptors. J Neurosci 35: 13133–13147, 2015. doi: 10.1523/JNEUROSCI.0658-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Grabner CP, Ratliff CP, Light AC, DeVries SH. Mechanism of High-Frequency Signaling at a Depressing Ribbon Synapse. Neuron 91: 133–145, 2016. doi: 10.1016/j.neuron.2016.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Grant L, Yi E, Glowatzki E. Two modes of release shape the postsynaptic response at the inner hair cell ribbon synapse. J Neurosci 30: 4210–4220, 2010. doi: 10.1523/JNEUROSCI.4439-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Grassmeyer JJ, Thoreson WB. Synaptic Ribbon Active Zones in Cone Photoreceptors Operate Independently from One Another. Front Cell Neurosci 11: 198, 2017. doi: 10.3389/fncel.2017.00198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Grauel MK, Maglione M, Reddy-Alla S, Willmes CG, Brockmann MM, Trimbuch T, Rosenmund T, Pangalos M, Vardar G, Stumpf A, Walter AM, Rost BR, Eickholt BJ, Haucke V, Schmitz D, Sigrist SJ, Rosenmund C. RIM-binding protein 2 regulates release probability by fine-tuning calcium channel localization at murine hippocampal synapses. Proc Natl Acad Sci USA 113: 11615–11620, 2016. doi: 10.1073/pnas.1605256113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Graydon CW, Cho S, Li G-L, Kachar B, von Gersdorff H. Sharp Ca2+ nanodomains beneath the ribbon promote highly synchronous multivesicular release at hair cell synapses. J Neurosci 31: 16637–16650, 2011. doi: 10.1523/JNEUROSCI.1866-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Greengard P, Valtorta F, Czernik AJ, Benfenati F. Synaptic vesicle phosphoproteins and regulation of synaptic function. Science 259: 780–785, 1993. doi: 10.1126/science.8430330. [DOI] [PubMed] [Google Scholar]
  • 131.Gregory FD, Bryan KE, Pangršič T, Calin-Jageman IE, Moser T, Lee A. Harmonin inhibits presynaptic Cav1.3 Ca2+ channels in mouse inner hair cells. Nat Neurosci 14: 1109–1111, 2011. doi: 10.1038/nn.2895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Gregory FD, Pangršič T, Calin-Jageman IE, Moser T, Lee A. Harmonin enhances voltage-dependent facilitation of Cav 1.3 channels and synchronous exocytosis in mouse inner hair cells: harmonin regulation of Cav 1.3 and exocytosis. J Physiol 591: 3253–3269, 2013. doi: 10.1113/jphysiol.2013.254367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Grillet N, Xiong W, Reynolds A, Kazmierczak P, Sato T, Lillo C, Dumont RA, Hintermann E, Sczaniecka A, Schwander M, Williams D, Kachar B, Gillespie PG, Müller U. Harmonin mutations cause mechanotransduction defects in cochlear hair cells. Neuron 62: 375–387, 2009. doi: 10.1016/j.neuron.2009.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Guillet M, Sendin G, Bourien J, Puel J-L, Nouvian R. Actin Filaments Regulate Exocytosis at the Hair Cell Ribbon Synapse. J Neurosci 36: 649–654, 2016. doi: 10.1523/JNEUROSCI.3379-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Gundelfinger ED, Reissner C, Garner CC. Role of Bassoon and Piccolo in Assembly and Molecular Organization of the Active Zone. Front Synaptic Neurosci 7: 19, 2016. doi: 10.3389/fnsyn.2015.00019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Haeseleer F, Imanishi Y, Maeda T, Possin DE, Maeda A, Lee A, Rieke F, Palczewski K. Essential role of Ca2+-binding protein 4, a Cav1.4 channel regulator, in photoreceptor synaptic function. Nat Neurosci 7: 1079–1087, 2004. doi: 10.1038/nn1320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Haeseleer F, Imanishi Y, Sokal I, Filipek S, Palczewski K. Calcium-binding proteins: intracellular sensors from the calmodulin superfamily. Biochem Biophys Res Commun 290: 615–623, 2002. doi: 10.1006/bbrc.2001.6228. [DOI] [PubMed] [Google Scholar]
  • 138.Haeseleer F, Sokal I, Verlinde CL, Erdjument-Bromage H, Tempst P, Pronin AN, Benovic JL, Fariss RN, Palczewski K. Five members of a novel Ca(2+)-binding protein (CABP) subfamily with similarity to calmodulin. J Biol Chem 275: 1247–1260, 2000. doi: 10.1074/jbc.275.2.1247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Haeseleer F, Williams B, Lee A. Characterization of C-terminal Splice Variants of Cav1.4 Ca2+ Channels in Human Retina. J Biol Chem 291: 15663–15673, 2016. doi: 10.1074/jbc.M116.731737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Han Y, Kaeser PS, Südhof TC, Schneggenburger R. RIM determines Ca2+ channel density and vesicle docking at the presynaptic active zone. Neuron 69: 304–316, 2011. doi: 10.1016/j.neuron.2010.12.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Hardie J, Lee A. Decalmodulation of Cav1 channels by CaBPs. Channels (Austin) 10: 33–37, 2016. doi: 10.1080/19336950.2015.1051273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Heidelberger R. Adenosine triphosphate and the late steps in calcium-dependent exocytosis at a ribbon synapse. J Gen Physiol 111: 225–241, 1998. doi: 10.1085/jgp.111.2.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Heidelberger R, Heinemann C, Neher E, Matthews G. Calcium dependence of the rate of exocytosis in a synaptic terminal. Nature 371: 513–515, 1994. doi: 10.1038/371513a0. [DOI] [PubMed] [Google Scholar]
  • 144.Heidelberger R, Wang MM, Sherry DM. Differential distribution of synaptotagmin immunoreactivity among synapses in the goldfish, salamander, and mouse retina. Vis Neurosci 20: 37–49, 2003. doi: 10.1017/S095252380320105X. [DOI] [PubMed] [Google Scholar]
  • 145.Heinemann C, Chow RH, Neher E, Zucker RS. Kinetics of the secretory response in bovine chromaffin cells following flash photolysis of caged Ca2+. Biophys J 67: 2546–2557, 1994. doi: 10.1016/S0006-3495(94)80744-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Hemara-Wahanui A, Berjukow S, Hope CI, Dearden PK, Wu S-B, Wilson-Wheeler J, Sharp DM, Lundon-Treweek P, Clover GM, Hoda J-C, Striessnig J, Marksteiner R, Hering S, Maw MA. A CACNA1F mutation identified in an X-linked retinal disorder shifts the voltage dependence of Cav1.4 channel activation. Proc Natl Acad Sci USA 102: 7553–7558, 2005. doi: 10.1073/pnas.0501907102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Hibino H, Pironkova R, Onwumere O, Vologodskaia M, Hudspeth AJ, Lesage F. RIM binding proteins (RBPs) couple Rab3-interacting molecules (RIMs) to voltage-gated Ca(2+) channels. Neuron 34: 411–423, 2002. doi: 10.1016/S0896-6273(02)00667-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Highstein SM, Holstein GR, Mann MA, Rabbitt RD. Evidence that protons act as neurotransmitters at vestibular hair cell-calyx afferent synapses. Proc Natl Acad Sci USA 111: 5421–5426, 2014. doi: 10.1073/pnas.1319561111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Hirasawa H, Yamada M, Kaneko A. Acidification of the synaptic cleft of cone photoreceptor terminal controls the amount of transmitter release, thereby forming the receptive field surround in the vertebrate retina. J Physiol Sci 62: 359–375, 2012. doi: 10.1007/s12576-012-0220-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Hoda J-C, Zaghetto F, Koschak A, Striessnig J. Congenital stationary night blindness type 2 mutations S229P, G369D, L1068P, and W1440X alter channel gating or functional expression of Ca(v)1.4 L-type Ca2+ channels. J Neurosci 25: 252–259, 2005. doi: 10.1523/JNEUROSCI.3054-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Hoda J-C, Zaghetto F, Singh A, Koschak A, Striessnig J. Effects of congenital stationary night blindness type 2 mutations R508Q and L1364H on Cav1.4 L-type Ca2+ channel function and expression. J Neurochem 96: 1648–1658, 2006. doi: 10.1111/j.1471-4159.2006.03678.x. [DOI] [PubMed] [Google Scholar]
  • 152.Hope CI, Sharp DM, Hemara-Wahanui A, Sissingh JI, Lundon P, Mitchell EA, Maw MA, Clover GM. Clinical manifestations of a unique X-linked retinal disorder in a large New Zealand family with a novel mutation in CACNA1F, the gene responsible for CSNB2. Clin Experiment Ophthalmol 33: 129–136, 2005. doi: 10.1111/j.1442-9071.2005.00987.x. [DOI] [PubMed] [Google Scholar]
  • 153.Hoppa MB, Lana B, Margas W, Dolphin AC, Ryan TA. α2δ expression sets presynaptic calcium channel abundance and release probability. Nature 486: 122–125, 2012. doi: 10.1038/nature11033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Hosoi N, Arai I, Tachibana M. Group III metabotropic glutamate receptors and exocytosed protons inhibit L-type calcium currents in cones but not in rods. J Neurosci 25: 4062–4072, 2005. doi: 10.1523/JNEUROSCI.2735-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Hu Z, Wang J-W, Yu D, Soon JL, de Kleijn DPV, Foo R, Liao P, Colecraft HM, Soong TW. Aberrant Splicing Promotes Proteasomal Degradation of L-type CaV1.2 Calcium Channels by Competitive Binding for CaVβ Subunits in Cardiac Hypertrophy. Sci Rep 6: 35247, 2016. doi: 10.1038/srep35247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Hulme JT, Yarov-Yarovoy V, Lin TW-C, Scheuer T, Catterall WA. Autoinhibitory control of the CaV1.2 channel by its proteolytically processed distal C-terminal domain. J Physiol 576: 87–102, 2006. doi: 10.1113/jphysiol.2006.111799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Isaacson JS, Walmsley B. Counting quanta: direct measurements of transmitter release at a central synapse. Neuron 15: 875–884, 1995. doi: 10.1016/0896-6273(95)90178-7. [DOI] [PubMed] [Google Scholar]
  • 158.Issa NP, Hudspeth AJ. The entry and clearance of Ca2+ at individual presynaptic active zones of hair cells from the bullfrog’s sacculus. Proc Natl Acad Sci USA 93: 9527–9532, 1996. doi: 10.1073/pnas.93.18.9527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Jackman SL, Choi S-Y, Thoreson WB, Rabl K, Bartoletti TM, Kramer RH. Role of the synaptic ribbon in transmitting the cone light response. Nat Neurosci 12: 303–310, 2009. doi: 10.1038/nn.2267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Jarsky T, Tian M, Singer JH. Nanodomain control of exocytosis is responsible for the signaling capability of a retinal ribbon synapse. J Neurosci 30: 11885–11895, 2010. doi: 10.1523/JNEUROSCI.1415-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Jean P, Lopez de la Morena D, Michanski S, Jaime Tobón LM, Chakrabarti R, Picher MM, Neef J, Jung S, Gültas M, Maxeiner S, Neef A, Wichmann C, Strenzke N, Grabner C, Moser T. The synaptic ribbon is critical for sound encoding at high rates and with temporal precision. eLife 7: e29275, 2018. doi: 10.7554/eLife.29275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Jian K, Barhoumi R, Ko ML, Ko GY-P. Inhibitory effect of somatostatin-14 on L-type voltage-gated calcium channels in cultured cone photoreceptors requires intracellular calcium. J Neurophysiol 102: 1801–1810, 2009. doi: 10.1152/jn.00354.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Jing Z, Rutherford MA, Takago H, Frank T, Fejtova A, Khimich D, Moser T, Strenzke N. Disruption of the presynaptic cytomatrix protein bassoon degrades ribbon anchorage, multiquantal release, and sound encoding at the hair cell afferent synapse. J Neurosci 33: 4456–4467, 2013. doi: 10.1523/JNEUROSCI.3491-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Johnson KR, Gagnon LH, Webb LS, Peters LL, Hawes NL, Chang B, Zheng QY. Mouse models of USH1C and DFNB18: phenotypic and molecular analyses of two new spontaneous mutations of the Ush1c gene. Hum Mol Genet 12: 3075–3086, 2003. doi: 10.1093/hmg/ddg332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Johnson SL, Ceriani F, Houston O, Polishchuk R, Polishchuk E, Crispino G, Zorzi V, Mammano F, Marcotti W. Connexin-Mediated Signaling in Nonsensory Cells Is Crucial for the Development of Sensory Inner Hair Cells in the Mouse Cochlea. J Neurosci 37: 258–268, 2017. doi: 10.1523/JNEUROSCI.2251-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Johnson SL, Eckrich T, Kuhn S, Zampini V, Franz C, Ranatunga KM, Roberts TP, Masetto S, Knipper M, Kros CJ, Marcotti W. Position-dependent patterning of spontaneous action potentials in immature cochlear inner hair cells. Nat Neurosci 14: 711–717, 2011. doi: 10.1038/nn.2803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Johnson SL, Franz C, Knipper M, Marcotti W. Functional maturation of the exocytotic machinery at gerbil hair cell ribbon synapses. J Physiol 587: 1715–1726, 2009. doi: 10.1113/jphysiol.2009.168542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Johnson SL, Marcotti W. Biophysical properties of CaV1.3 calcium channels in gerbil inner hair cells. J Physiol 586: 1029–1042, 2008. doi: 10.1113/jphysiol.2007.145219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Johnson SL, Olt J, Cho S, von Gersdorff H, Marcotti W. The Coupling between Ca2+ Channels and the Exocytotic Ca2+ Sensor at Hair Cell Ribbon Synapses Varies Tonotopically along the Mature Cochlea. J Neurosci 37: 2471–2484, 2017. doi: 10.1523/JNEUROSCI.2867-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Jung S, Oshima-Takago T, Chakrabarti R, Wong AB, Jing Z, Yamanbaeva G, Picher MM, Wojcik SM, Göttfert F, Predoehl F, Michel K, Hell SW, Schoch S, Strenzke N, Wichmann C, Moser T. Rab3-interacting molecules 2α and 2β promote the abundance of voltage-gated CaV1.3 Ca2+ channels at hair cell active zones. Proc Natl Acad Sci USA 112: E3141–E3149, 2015. doi: 10.1073/pnas.1417207112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Jurkat-Rott K, Lehmann-Horn F. The impact of splice isoforms on voltage-gated calcium channel alpha1 subunits. J Physiol 554: 609–619, 2004. doi: 10.1113/jphysiol.2003.052712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Kaeser PS, Deng L, Wang Y, Dulubova I, Liu X, Rizo J, Südhof TC. RIM proteins tether Ca2+ channels to presynaptic active zones via a direct PDZ-domain interaction. Cell 144: 282–295, 2011. doi: 10.1016/j.cell.2010.12.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Kandler K, Clause A, Noh J. Tonotopic reorganization of developing auditory brainstem circuits. Nat Neurosci 12: 711–717, 2009. doi: 10.1038/nn.2332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Katoh M. GIPC gene family (Review). Int J Mol Med 9: 585–589, 2002. [PubMed] [Google Scholar]
  • 175.Katz B, Miledi R. A study of synaptic transmission in the absence of nerve impulses. J Physiol 192: 407–436, 1967. doi: 10.1113/jphysiol.1967.sp008307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Keen EC, Hudspeth AJ. Transfer characteristics of the hair cell’s afferent synapse. Proc Natl Acad Sci USA 103: 5537–5542, 2006. doi: 10.1073/pnas.0601103103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Kersten FFJ, van Wijk E, van Reeuwijk J, van der Zwaag B, Märker T, Peters TA, Katsanis N, Wolfrum U, Keunen JEE, Roepman R, Kremer H. Association of whirlin with Cav1.3 (alpha1D) channels in photoreceptors, defining a novel member of the usher protein network. Invest Ophthalmol Vis Sci 51: 2338–2346, 2010. doi: 10.1167/iovs.09-4650. [DOI] [PubMed] [Google Scholar]
  • 178.Khimich D, Nouvian R, Pujol R, Tom Dieck S, Egner A, Gundelfinger ED, Moser T. Hair cell synaptic ribbons are essential for synchronous auditory signalling. Nature 434: 889–894, 2005. doi: 10.1038/nature03418. [DOI] [PubMed] [Google Scholar]
  • 179.Kim EY, Rumpf CH, Van Petegem F, Arant RJ, Findeisen F, Cooley ES, Isacoff EY, Minor DL Jr. Multiple C-terminal tail Ca(2+)/CaMs regulate Ca(V)1.2 function but do not mediate channel dimerization. EMBO J 29: 3924–3938, 2010. doi: 10.1038/emboj.2010.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Kim KY, Scholl ES, Liu X, Shepherd A, Haeseleer F, Lee A. Localization and expression of CaBP1/caldendrin in the mouse brain. Neuroscience 268: 33–47, 2014. doi: 10.1016/j.neuroscience.2014.02.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Kiyonaka S, Nakajima H, Takada Y, Hida Y, Yoshioka T, Hagiwara A, Kitajima I, Mori Y, Ohtsuka T. Physical and functional interaction of the active zone protein CAST/ERC2 and the β-subunit of the voltage-dependent Ca(2+) channel. J Biochem 152: 149–159, 2012. doi: 10.1093/jb/mvs054. [DOI] [PubMed] [Google Scholar]
  • 182.Kiyonaka S, Wakamori M, Miki T, Uriu Y, Nonaka M, Bito H, Beedle AM, Mori E, Hara Y, De Waard M, Kanagawa M, Itakura M, Takahashi M, Campbell KP, Mori Y. RIM1 confers sustained activity and neurotransmitter vesicle anchoring to presynaptic Ca2+ channels. Nat Neurosci 10: 691–701, 2007. doi: 10.1038/nn1904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Klöckner U, Isenberg G. Calcium channel current of vascular smooth muscle cells: extracellular protons modulate gating and single channel conductance. J Gen Physiol 103: 665–678, 1994. doi: 10.1085/jgp.103.4.665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Knirsch M, Brandt N, Braig C, Kuhn S, Hirt B, Münkner S, Knipper M, Engel J. Persistence of Ca(v)1.3 Ca2+ channels in mature outer hair cells supports outer hair cell afferent signaling. J Neurosci 27: 6442–6451, 2007. doi: 10.1523/JNEUROSCI.5364-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Knoflach D, Kerov V, Sartori SB, Obermair GJ, Schmuckermair C, Liu X, Sothilingam V, Garcia Garrido M, Baker SA, Glösmann M, Schicker K, Seeliger M, Lee A, Koschak A. Cav1.4 IT mouse as model for vision impairment in human congenital stationary night blindness type 2. Channels (Austin) 7: 503–513, 2013. doi: 10.4161/chan.26368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Knoflach D, Schicker K, Glösmann M, Koschak A. Gain-of-function nature of Cav1.4 L-type calcium channels alters firing properties of mouse retinal ganglion cells. Channels (Austin) 9: 298–306, 2015. doi: 10.1080/19336950.2015.1078040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Kolb H, Dekorver L. Midget ganglion cells of the parafovea of the human retina: a study by electron microscopy and serial section reconstructions. J Comp Neurol 303: 617–636, 1991. doi: 10.1002/cne.903030408. [DOI] [PubMed] [Google Scholar]
  • 188.Kollmar R, Fak J, Montgomery LG, Hudspeth AJ. Hair cell-specific splicing of mRNA for the alpha1D subunit of voltage-gated Ca2+ channels in the chicken’s cochlea. Proc Natl Acad Sci USA 94: 14889–14893, 1997. doi: 10.1073/pnas.94.26.14889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Kollmar R, Montgomery LG, Fak J, Henry LJ, Hudspeth AJ. Predominance of the alpha1D subunit in L-type voltage-gated Ca2+ channels of hair cells in the chicken’s cochlea. Proc Natl Acad Sci USA 94: 14883–14888, 1997. doi: 10.1073/pnas.94.26.14883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Koschak A, Reimer D, Huber I, Grabner M, Glossmann H, Engel J, Striessnig J. alpha 1D (Cav1.3) subunits can form l-type Ca2+ channels activating at negative voltages. J Biol Chem 276: 22100–22106, 2001. doi: 10.1074/jbc.M101469200. [DOI] [PubMed] [Google Scholar]
  • 191.Koschak A, Reimer D, Walter D, Hoda J-C, Heinzle T, Grabner M, Striessnig J. Cav1.4alpha1 subunits can form slowly inactivating dihydropyridine-sensitive L-type Ca2+ channels lacking Ca2+-dependent inactivation. J Neurosci 23: 6041–6049, 2003. doi: 10.1523/JNEUROSCI.23-14-06041.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Kourennyi DE, Barnes S. Depolarization-induced calcium channel facilitation in rod photoreceptors is independent of G proteins and phosphorylation. J Neurophysiol 84: 133–138, 2000. doi: 10.1152/jn.2000.84.1.133. [DOI] [PubMed] [Google Scholar]
  • 193.Kourennyi DE, Liu XD, Hart J, Mahmud F, Baldridge WH, Barnes S. Reciprocal modulation of calcium dynamics at rod and cone photoreceptor synapses by nitric oxide. J Neurophysiol 92: 477–483, 2004. doi: 10.1152/jn.00606.2003. [DOI] [PubMed] [Google Scholar]
  • 194.Krafte DS, Kass RS. Hydrogen ion modulation of Ca channel current in cardiac ventricular cells. Evidence for multiple mechanisms. J Gen Physiol 91: 641–657, 1988. doi: 10.1085/jgp.91.5.641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Krinner S, Butola T, Jung S, Wichmann C, Moser T. RIM-binding protein 2 promotes a large number of CaV1.3 Ca2+-channels and contributes to fast synaptic vesicle replenishment at hair cell active zones. Front Cell Neurosci 11: 334, 2017. doi: 10.3389/fncel.2017.00334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Kros CJ, Ruppersberg JP, Rüsch A. Expression of a potassium current in inner hair cells during development of hearing in mice. Nature 394: 281–284, 1998. doi: 10.1038/28401. [DOI] [PubMed] [Google Scholar]
  • 197.Kuhn S, Knirsch M, Rüttiger L, Kasperek S, Winter H, Freichel M, Flockerzi V, Knipper M, Engel J. Ba2+ currents in inner and outer hair cells of mice lacking the voltage-dependent Ca2+ channel subunits beta3 or beta4. Channels (Austin) 3: 366–376, 2009. doi: 10.4161/chan.3.5.9774. [DOI] [PubMed] [Google Scholar]
  • 198.Kujawa SG, Liberman MC. Adding insult to injury: cochlear nerve degeneration after “temporary” noise-induced hearing loss. J Neurosci 29: 14077–14085, 2009. doi: 10.1523/JNEUROSCI.2845-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Lee A, Wang S, Williams B, Hagen J, Scheetz TE, Haeseleer F. Characterization of Cav1.4 complexes (α11.4, β2, and α2δ4) in HEK293T cells and in the retina. J Biol Chem 290: 1505–1521, 2015. doi: 10.1074/jbc.M114.607465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Lentz JJ, Gordon WC, Farris HE, MacDonald GH, Cunningham DE, Robbins CA, Tempel BL, Bazan NG, Rubel EW, Oesterle EC, Keats BJ. Deafness and retinal degeneration in a novel USH1C knock-in mouse model. Dev Neurobiol 70: 253–267, 2010. doi: 10.1002/dneu.20771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Levelt CN, Hübener M. Critical-period plasticity in the visual cortex. Annu Rev Neurosci 35: 309–330, 2012. doi: 10.1146/annurev-neuro-061010-113813. [DOI] [PubMed] [Google Scholar]
  • 202.Levic S, Dulon D. The temporal characteristics of Ca2+ entry through L-type and T-type Ca2+ channels shape exocytosis efficiency in chick auditory hair cells during development. J Neurophysiol 108: 3116–3123, 2012. doi: 10.1152/jn.00555.2012. [DOI] [PubMed] [Google Scholar]
  • 203.Levic S, Nie L, Tuteja D, Harvey M, Sokolowski BHA, Yamoah EN. Development and regeneration of hair cells share common functional features. Proc Natl Acad Sci USA 104: 19108–19113, 2007. doi: 10.1073/pnas.0705927104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Li G-L, Cho S, von Gersdorff H. Phase-locking precision is enhanced by multiquantal release at an auditory hair cell ribbon synapse. Neuron 83: 1404–1417, 2014. doi: 10.1016/j.neuron.2014.08.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Li G-L, Keen E, Andor-Ardó D, Hudspeth AJ, von Gersdorff H. The unitary event underlying multiquantal EPSCs at a hair cell’s ribbon synapse. J Neurosci 29: 7558–7568, 2009. doi: 10.1523/JNEUROSCI.0514-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Liberman LD, Liberman MC. Dynamics of cochlear synaptopathy after acoustic overexposure. J Assoc Res Otolaryngol 16: 205–219, 2015. doi: 10.1007/s10162-015-0510-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Liberman LD, Wang H, Liberman MC. Opposing gradients of ribbon size and AMPA receptor expression underlie sensitivity differences among cochlear-nerve/hair-cell synapses. J Neurosci 31: 801–808, 2011. doi: 10.1523/JNEUROSCI.3389-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Liberman MC. Single-neuron labeling in the cat auditory nerve. Science 216: 1239–1241, 1982. doi: 10.1126/science.7079757. [DOI] [PubMed] [Google Scholar]
  • 209.Liberman MC. Auditory-nerve response from cats raised in a low-noise chamber. J Acoust Soc Am 63: 442–455, 1978. doi: 10.1121/1.381736. [DOI] [PubMed] [Google Scholar]
  • 210.Liberman MC. Morphological differences among radial afferent fibers in the cat cochlea: an electron-microscopic study of serial sections. Hear Res 3: 45–63, 1980. doi: 10.1016/0378-5955(80)90007-6. [DOI] [PubMed] [Google Scholar]
  • 211.Lieb A, Ortner N, Striessnig J. C-terminal modulatory domain controls coupling of voltage-sensing to pore opening in Cav1.3 L-type Ca(2+) channels. Biophys J 106: 1467–1475, 2014. doi: 10.1016/j.bpj.2014.02.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Littink KW, van Genderen MM, Collin RWJ, Roosing S, de Brouwer APM, Riemslag FCC, Venselaar H, Thiadens AAHJ, Hoyng CB, Rohrschneider K, den Hollander AI, Cremers FPM, van den Born LI. A novel homozygous nonsense mutation in CABP4 causes congenital cone-rod synaptic disorder. Invest Ophthalmol Vis Sci 50: 2344–2350, 2009. doi: 10.1167/iovs.08-2553. [DOI] [PubMed] [Google Scholar]
  • 213.Liu C, Bickford LS, Held RG, Nyitrai H, Südhof TC, Kaeser PS. The active zone protein family ELKS supports Ca2+ influx at nerve terminals of inhibitory hippocampal neurons. J Neurosci 34: 12289–12303, 2014. doi: 10.1523/JNEUROSCI.0999-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Liu KSY, Siebert M, Mertel S, Knoche E, Wegener S, Wichmann C, Matkovic T, Muhammad K, Depner H, Mettke C, Bückers J, Hell SW, Müller M, Davis GW, Schmitz D, Sigrist SJ. RIM-binding protein, a central part of the active zone, is essential for neurotransmitter release. Science 334: 1565–1569, 2011. doi: 10.1126/science.1212991. [DOI] [PubMed] [Google Scholar]
  • 215.Liu X, Kerov V, Haeseleer F, Majumder A, Artemyev N, Baker SA, Lee A. Dysregulation of Ca(v)1.4 channels disrupts the maturation of photoreceptor synaptic ribbons in congenital stationary night blindness type 2. Channels (Austin) 7: 514–523, 2013. doi: 10.4161/chan.26376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Liu X, Yang PS, Yang W, Yue DT. Enzyme-inhibitor-like tuning of Ca(2+) channel connectivity with calmodulin. Nature 463: 968–972, 2010. doi: 10.1038/nature08766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Llinás R, Steinberg IZ, Walton K. Relationship between presynaptic calcium current and postsynaptic potential in squid giant synapse. Biophys J 33: 323–351, 1981. doi: 10.1016/S0006-3495(81)84899-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Löhner M, Babai N, Müller T, Gierke K, Atorf J, Joachimsthaler A, Peukert A, Martens H, Feigenspan A, Kremers J, Schoch S, Brandstätter JH, Regus-Leidig H. Analysis of RIM Expression and Function at Mouse Photoreceptor Ribbon Synapses. J Neurosci 37: 7848–7863, 2017. doi: 10.1523/JNEUROSCI.2795-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Lu L, Sirish P, Zhang Z, Woltz RL, Li N, Timofeyev V, Knowlton AA, Zhang X-D, Yamoah EN, Chiamvimonvat N. Regulation of gene transcription by voltage-gated L-type calcium channel, Cav1.3. J Biol Chem 290: 4663–4676, 2015. doi: 10.1074/jbc.M114.586883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Lv C, Gould TJ, Bewersdorf J, Zenisek D. High-resolution optical imaging of zebrafish larval ribbon synapse protein RIBEYE, RIM2, and CaV 1.4 by stimulation emission depletion microscopy. Microsc Microanal 18: 745–752, 2012. doi: 10.1017/S1431927612000268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Lv C, Stewart WJ, Akanyeti O, Frederick C, Zhu J, Santos-Sacchi J, Sheets L, Liao JC, Zenisek D. Synaptic Ribbons Require Ribeye for Electron Density, Proper Synaptic Localization, and Recruitment of Calcium Channels. Cell Reports 15: 2784–2795, 2016. doi: 10.1016/j.celrep.2016.05.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Magupalli VG, Schwarz K, Alpadi K, Natarajan S, Seigel GM, Schmitz F. Multiple RIBEYE-RIBEYE interactions create a dynamic scaffold for the formation of synaptic ribbons. J Neurosci 28: 7954–7967, 2008. doi: 10.1523/JNEUROSCI.1964-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Mandell JW, Townes-Anderson E, Czernik AJ, Cameron R, Greengard P, De Camilli P. Synapsins in the vertebrate retina: absence from ribbon synapses and heterogeneous distribution among conventional synapses. Neuron 5: 19–33, 1990. doi: 10.1016/0896-6273(90)90030-J. [DOI] [PubMed] [Google Scholar]
  • 224.Mangoni ME, Couette B, Bourinet E, Platzer J, Reimer D, Striessnig J, Nargeot J. Functional role of L-type Cav1.3 Ca2+ channels in cardiac pacemaker activity. Proc Natl Acad Sci USA 100: 5543–5548, 2003. doi: 10.1073/pnas.0935295100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Mansergh F, Orton NC, Vessey JP, Lalonde MR, Stell WK, Tremblay F, Barnes S, Rancourt DE, Bech-Hansen NT. Mutation of the calcium channel gene Cacna1f disrupts calcium signaling, synaptic transmission and cellular organization in mouse retina. Hum Mol Genet 14: 3035–3046, 2005. doi: 10.1093/hmg/ddi336. [DOI] [PubMed] [Google Scholar]
  • 226.Marangoudakis S, Andrade A, Helton TD, Denome S, Castiglioni AJ, Lipscombe D. Differential ubiquitination and proteasome regulation of Ca(V)2.2 N-type channel splice isoforms. J Neurosci 32: 10365–10369, 2012. doi: 10.1523/JNEUROSCI.0851-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Marcotti W, Johnson SL, Kros CJ. A transiently expressed SK current sustains and modulates action potential activity in immature mouse inner hair cells. J Physiol 560: 691–708, 2004. doi: 10.1113/jphysiol.2004.072868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Marcotti W, Johnson SL, Holley MC, Kros CJ. Developmental changes in the expression of potassium currents of embryonic, neonatal and mature mouse inner hair cells. J Physiol 548: 383–400, 2003. doi: 10.1113/jphysiol.2002.034801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Marcotti W, Johnson SL, Rüsch A, Kros CJ. Sodium and calcium currents shape action potentials in immature mouse inner hair cells. J Physiol 552: 743–761, 2003. doi: 10.1113/jphysiol.2003.043612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Martini M, Rossi ML, Rubbini G, Rispoli G. Calcium currents in hair cells isolated from semicircular canals of the frog. Biophys J 78: 1240–1254, 2000. doi: 10.1016/S0006-3495(00)76681-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Masland RH. The neuronal organization of the retina. Neuron 76: 266–280, 2012. doi: 10.1016/j.neuron.2012.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Matveev V, Bertram R, Sherman A. Calcium cooperativity of exocytosis as a measure of Ca2+ channel domain overlap. Brain Res 1398: 126–138, 2011. doi: 10.1016/j.brainres.2011.05.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Maxeiner S, Luo F, Tan A, Schmitz F, Südhof TC. How to make a synaptic ribbon: RIBEYE deletion abolishes ribbons in retinal synapses and disrupts neurotransmitter release. EMBO J 35: 1098–1114, 2016. doi: 10.15252/embj.201592701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Maximov A, Bezprozvanny I. Synaptic targeting of N-type calcium channels in hippocampal neurons. J Neurosci 22: 6939–6952, 2002. doi: 10.1523/JNEUROSCI.22-16-06939.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.McLean WJ, Smith KA, Glowatzki E, Pyott SJ. Distribution of the Na,K-ATPase α subunit in the rat spiral ganglion and organ of Corti. J Assoc Res Otolaryngol 10: 37–49, 2009. doi: 10.1007/s10162-008-0152-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.McRory JE, Hamid J, Doering CJ, Garcia E, Parker R, Hamming K, Chen L, Hildebrand M, Beedle AM, Feldcamp L, Zamponi GW, Snutch TP. The CACNA1F gene encodes an L-type calcium channel with unique biophysical properties and tissue distribution. J Neurosci 24: 1707–1718, 2004. doi: 10.1523/JNEUROSCI.4846-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Mehta B, Ke J-B, Zhang L, Baden AD, Markowitz AL, Nayak S, Briggman KL, Zenisek D, Singer JH. Global Ca2+ signaling drives ribbon-independent synaptic transmission at rod bipolar cell synapses. J Neurosci 34: 6233–6244, 2014. doi: 10.1523/JNEUROSCI.5324-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Mehta B, Snellman J, Chen S, Li W, Zenisek D. Synaptic ribbons influence the size and frequency of miniature-like evoked postsynaptic currents. Neuron 77: 516–527, 2013. doi: 10.1016/j.neuron.2012.11.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Meister M, Wong RO, Baylor DA, Shatz CJ. Synchronous bursts of action potentials in ganglion cells of the developing mammalian retina. Science 252: 939–943, 1991. doi: 10.1126/science.2035024. [DOI] [PubMed] [Google Scholar]
  • 240.Mercer AJ, Chen M, Thoreson WB. Lateral mobility of presynaptic L-type calcium channels at photoreceptor ribbon synapses. J Neurosci 31: 4397–4406, 2011. doi: 10.1523/JNEUROSCI.5921-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Mercer AJ, Szalewski RJ, Jackman SL, Van Hook MJ, Thoreson WB. Regulation of presynaptic strength by controlling Ca2+ channel mobility: effects of cholesterol depletion on release at the cone ribbon synapse. J Neurophysiol 107: 3468–3478, 2012. doi: 10.1152/jn.00779.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Mercer AJ, Thoreson WB. The dynamic architecture of photoreceptor ribbon synapses: cytoskeletal, extracellular matrix, and intramembrane proteins. Vis Neurosci 28: 453–471, 2011. doi: 10.1017/S0952523811000356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Merchan-Perez A, Liberman MC. Ultrastructural differences among afferent synapses on cochlear hair cells: correlations with spontaneous discharge rate. J Comp Neurol 371: 208–221, 1996. doi:. [DOI] [PubMed] [Google Scholar]
  • 244.Meyer AC, Frank T, Khimich D, Hoch G, Riedel D, Chapochnikov NM, Yarin YM, Harke B, Hell SW, Egner A, Moser T. Tuning of synapse number, structure and function in the cochlea. Nat Neurosci 12: 444–453, 2009. doi: 10.1038/nn.2293. [DOI] [PubMed] [Google Scholar]
  • 245.Michalski N, Michel V, Caberlotto E, Lefèvre GM, van Aken AFJ, Tinevez J-Y, Bizard E, Houbron C, Weil D, Hardelin J-P, Richardson GP, Kros CJ, Martin P, Petit C. Harmonin-b, an actin-binding scaffold protein, is involved in the adaptation of mechanoelectrical transduction by sensory hair cells. Pflugers Arch 459: 115–130, 2009. doi: 10.1007/s00424-009-0711-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Miyake Y. [Establishment of the concept of new clinical entities–complete and incomplete form of congenital stationary night blindness]. Nippon Ganka Gakkai Zasshi 106: 737–755, 2002. [PubMed] [Google Scholar]
  • 247.Mochida S, Westenbroek RE, Yokoyama CT, Zhong H, Myers SJ, Scheuer T, Itoh K, Catterall WA. Requirement for the synaptic protein interaction site for reconstitution of synaptic transmission by P/Q-type calcium channels. Proc Natl Acad Sci USA 100: 2819–2824, 2003. doi: 10.1073/pnas.262787699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Monaghan P, Osborne MP. Light-induced formation of dense-core vesicles in rod photoreceptors in retina of Xenopus laevis. Nature 257: 586–587, 1975. doi: 10.1038/257586a0. [DOI] [PubMed] [Google Scholar]
  • 249.Moncrief ND, Kretsinger RH, Goodman M. Evolution of EF-hand calcium-modulated proteins. I. Relationships based on amino acid sequences. J Mol Evol 30: 522–562, 1990. doi: 10.1007/BF02101108. [DOI] [PubMed] [Google Scholar]
  • 250.Morgans CW. Calcium channel heterogeneity among cone photoreceptors in the tree shrew retina. Eur J Neurosci 11: 2989–2993, 1999. doi: 10.1046/j.1460-9568.1999.00719.x. [DOI] [PubMed] [Google Scholar]
  • 251.Morgans CW, Bayley PR, Oesch NW, Ren G, Akileswaran L, Taylor WR. Photoreceptor calcium channels: insight from night blindness. Vis Neurosci 22: 561–568, 2005. doi: 10.1017/S0952523805225038. [DOI] [PubMed] [Google Scholar]
  • 252.Mortensen LS, Park SJH, Ke J-B, Cooper BH, Zhang L, Imig C, Löwel S, Reim K, Brose N, Demb JB, Rhee J-S, Singer JH. Complexin 3 Increases the Fidelity of Signaling in a Retinal Circuit by Regulating Exocytosis at Ribbon Synapses. Cell Reports 15: 2239–2250, 2016. doi: 10.1016/j.celrep.2016.05.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Moser T, Beutner D. Kinetics of exocytosis and endocytosis at the cochlear inner hair cell afferent synapse of the mouse. Proc Natl Acad Sci USA 97: 883–888, 2000. doi: 10.1073/pnas.97.2.883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Moser T, Neef A, Khimich D. Mechanisms underlying the temporal precision of sound coding at the inner hair cell ribbon synapse. J Physiol 576: 55–62, 2006. doi: 10.1113/jphysiol.2006.114835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Naraghi M, Neher E. Linearized buffered Ca2+ diffusion in microdomains and its implications for calculation of [Ca2+] at the mouth of a calcium channel. J Neurosci 17: 6961–6973, 1997. doi: 10.1523/JNEUROSCI.17-18-06961.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Neef J, Gehrt A, Bulankina AV, Meyer AC, Riedel D, Gregg RG, Strenzke N, Moser T. The Ca2+ channel subunit beta2 regulates Ca2+ channel abundance and function in inner hair cells and is required for hearing. J Neurosci 29: 10730–10740, 2009. doi: 10.1523/JNEUROSCI.1577-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Neef J, Jung S, Wong AB, Reuter K, Pangrsic T, Chakrabarti R, Kügler S, Lenz C, Nouvian R, Boumil RM, Frankel WN, Wichmann C, Moser T. Modes and regulation of endocytic membrane retrieval in mouse auditory hair cells. J Neurosci 34: 705–716, 2014. doi: 10.1523/JNEUROSCI.3313-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Neef J, Urban NT, Ohn TL, Frank T, Jean P, Hell SW, Willig KI, Moser T. Quantitative optical nanophysiology of Ca2+ signaling at inner hair cell active zones. Nat Commun 9: 290, 2018. doi: 10.1038/s41467-017-02612-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Nemzou N RM, Bulankina AV, Khimich D, Giese A, Moser T. Synaptic organization in cochlear inner hair cells deficient for the CaV1.3 (alpha1D) subunit of L-type Ca2+ channels. Neuroscience 141: 1849–1860, 2006. doi: 10.1016/j.neuroscience.2006.05.057. [DOI] [PubMed] [Google Scholar]
  • 260.Nie L, Zhu J, Gratton MA, Liao A, Mu KJ, Nonner W, Richardson GP, Yamoah EN. Molecular identity and functional properties of a novel T-type Ca2+ channel cloned from the sensory epithelia of the mouse inner ear. J Neurophysiol 100: 2287–2299, 2008. doi: 10.1152/jn.90707.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Nouvian R, Neef J, Bulankina AV, Reisinger E, Pangršič T, Frank T, Sikorra S, Brose N, Binz T, Moser T. Exocytosis at the hair cell ribbon synapse apparently operates without neuronal SNARE proteins. Nat Neurosci 14: 411–413, 2011. doi: 10.1038/nn.2774. [DOI] [PubMed] [Google Scholar]
  • 262.Ohn T-L, Rutherford MA, Jing Z, Jung S, Duque-Afonso CJ, Hoch G, Picher MM, Scharinger A, Strenzke N, Moser T. Hair cells use active zones with different voltage dependence of Ca2+ influx to decompose sounds into complementary neural codes. Proc Natl Acad Sci USA 113: E4716–E4725, 2016. doi: 10.1073/pnas.1605737113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Okawa H, Hoon M, Yoshimatsu T, Della Santina L, Wong ROL. Illuminating the multifaceted roles of neurotransmission in shaping neuronal circuitry. Neuron 83: 1303–1318, 2014. doi: 10.1016/j.neuron.2014.08.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Omori Y, Araki F, Chaya T, Kajimura N, Irie S, Terada K, Muranishi Y, Tsujii T, Ueno S, Koyasu T, Tamaki Y, Kondo M, Amano S, Furukawa T. Presynaptic dystroglycan-pikachurin complex regulates the proper synaptic connection between retinal photoreceptor and bipolar cells. J Neurosci 32: 6126–6137, 2012. doi: 10.1523/JNEUROSCI.0322-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Oz S, Benmocha A, Sasson Y, Sachyani D, Almagor L, Lee A, Hirsch JA, Dascal N. Competitive and non-competitive regulation of calcium-dependent inactivation in CaV1.2 L-type Ca2+ channels by calmodulin and Ca2+-binding protein 1. J Biol Chem 288: 12680–12691, 2013. doi: 10.1074/jbc.M113.460949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Oz S, Tsemakhovich V, Christel CJ, Lee A, Dascal N. CaBP1 regulates voltage-dependent inactivation and activation of Ca(V)1.2 (L-type) calcium channels. J Biol Chem 286: 13945–13953, 2011. doi: 10.1074/jbc.M110.198424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Page KM, Rothwell SW, Dolphin AC. The CaVβ subunit protects the I–II loop of the voltage-gated calcium channel, CaV2.2, from proteasomal degradation but not oligo-ubiquitination. J Biol Chem 291: 20402–20416, 2016. doi: 10.1074/jbc.M116.737270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Palczewski K, Polans AS, Baehr W, Ames JB. Ca(2+)-binding proteins in the retina: structure, function, and the etiology of human visual diseases. BioEssays 22: 337–350, 2000. doi:. [DOI] [PubMed] [Google Scholar]
  • 269.Palmer MJ, Hull C, Vigh J, von Gersdorff H. Synaptic cleft acidification and modulation of short-term depression by exocytosed protons in retinal bipolar cells. J Neurosci 23: 11332–11341, 2003. doi: 10.1523/JNEUROSCI.23-36-11332.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Pangršič T, Gabrielaitis M, Michanski S, Schwaller B, Wolf F, Strenzke N, Moser T. EF-hand protein Ca2+ buffers regulate Ca2+ influx and exocytosis in sensory hair cells. Proc Natl Acad Sci USA 112: E1028–E1037, 2015. doi: 10.1073/pnas.1416424112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Pangrsic T, Lasarow L, Reuter K, Takago H, Schwander M, Riedel D, Frank T, Tarantino LM, Bailey JS, Strenzke N, Brose N, Müller U, Reisinger E, Moser T. Hearing requires otoferlin-dependent efficient replenishment of synaptic vesicles in hair cells. Nat Neurosci 13: 869–876, 2010. doi: 10.1038/nn.2578. [DOI] [PubMed] [Google Scholar]
  • 272.Pangršič T, Reisinger E, Moser T. Otoferlin: a multi-C2 domain protein essential for hearing. Trends Neurosci 35: 671–680, 2012. doi: 10.1016/j.tins.2012.08.002. [DOI] [PubMed] [Google Scholar]
  • 273.Park S, Li C, Haeseleer F, Palczewski K, Ames JB. Structural insights into activation of the retinal L-type Ca2+ channel (Cav1.4) by Ca2+-binding protein 4 (CaBP4). J Biol Chem 289: 31262–31273, 2014. doi: 10.1074/jbc.M114.604439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Parsons TD, Sterling P. Synaptic ribbon. Conveyor belt or safety belt? Neuron 37: 379–382, 2003. doi: 10.1016/S0896-6273(03)00062-X. [DOI] [PubMed] [Google Scholar]
  • 275.Peloquin JB, Rehak R, Doering CJ, McRory JE. Functional analysis of congenital stationary night blindness type-2 CACNA1F mutations F742C, G1007R, and R1049W. Neuroscience 150: 335–345, 2007. doi: 10.1016/j.neuroscience.2007.09.021. [DOI] [PubMed] [Google Scholar]
  • 276.Pérez de Sevilla Müller L, Sargoy A, Fernández-Sánchez L, Rodriguez A, Liu J, Cuenca N, Brecha N. Expression and cellular localization of the voltage-gated calcium channel α2δ3 in the rodent retina. J Comp Neurol 523: 1443–1460, 2015. doi: 10.1002/cne.23751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Picher MM, Gehrt A, Meese S, Ivanovic A, Predoehl F, Jung S, Schrauwen I, Dragonetti AG, Colombo R, Van Camp G, Strenzke N, Moser T. Ca2+-binding protein 2 inhibits Ca2+-channel inactivation in mouse inner hair cells. Proc Natl Acad Sci USA 114: E1717–E1726, 2017. doi: 10.1073/pnas.1617533114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Picher MM, Oprişoreanu A-M, Jung S, Michel K, Schoch S, Moser T. Rab Interacting Molecules 2 and 3 Directly Interact with the Pore-Forming CaV1.3 Ca2+ Channel Subunit and Promote Its Membrane Expression. Front Cell Neurosci 11: 160, 2017. doi: 10.3389/fncel.2017.00160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Pirone A, Kurt S, Zuccotti A, Rüttiger L, Pilz P, Brown DH, Franz C, Schweizer M, Rust MB, Rübsamen R, Friauf E, Knipper M, Engel J. α2δ3 is essential for normal structure and function of auditory nerve synapses and is a novel candidate for auditory processing disorders. J Neurosci 34: 434–445, 2014. doi: 10.1523/JNEUROSCI.3085-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Platzer J, Engel J, Schrott-Fischer A, Stephan K, Bova S, Chen H, Zheng H, Striessnig J. Congenital deafness and sinoatrial node dysfunction in mice lacking class D L-type Ca2+ channels. Cell 102: 89–97, 2000. doi: 10.1016/S0092-8674(00)00013-1. [DOI] [PubMed] [Google Scholar]
  • 281.Prod’hom B, Pietrobon D, Hess P. Direct measurement of proton transfer rates to a group controlling the dihydropyridine-sensitive Ca2+ channel. Nature 329: 243–246, 1987. doi: 10.1038/329243a0. [DOI] [PubMed] [Google Scholar]
  • 282.Pumplin DW, Reese TS, Llinás R. Are the presynaptic membrane particles the calcium channels? Proc Natl Acad Sci USA 78: 7210–7213, 1981. doi: 10.1073/pnas.78.11.7210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Rao R, Buchsbaum G, Sterling P. Rate of quantal transmitter release at the mammalian rod synapse. Biophys J 67: 57–63, 1994. doi: 10.1016/S0006-3495(94)80454-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Rao-Mirotznik R, Buchsbaum G, Sterling P. Transmitter concentration at a three-dimensional synapse. J Neurophysiol 80: 3163–3172, 1998. doi: 10.1152/jn.1998.80.6.3163. [DOI] [PubMed] [Google Scholar]
  • 285.Rao-Mirotznik R, Harkins AB, Buchsbaum G, Sterling P. Mammalian rod terminal: architecture of a binary synapse. Neuron 14: 561–569, 1995. doi: 10.1016/0896-6273(95)90312-7. [DOI] [PubMed] [Google Scholar]
  • 286.Raven MA, Orton NC, Nassar H, Williams GA, Stell WK, Jacobs GH, Bech-Hansen NT, Reese BE. Early afferent signaling in the outer plexiform layer regulates development of horizontal cell morphology. J Comp Neurol 506: 745–758, 2008. doi: 10.1002/cne.21526. [DOI] [PubMed] [Google Scholar]
  • 287.Rea R, Li J, Dharia A, Levitan ES, Sterling P, Kramer RH. Streamlined synaptic vesicle cycle in cone photoreceptor terminals. Neuron 41: 755–766, 2004. doi: 10.1016/S0896-6273(04)00088-1. [DOI] [PubMed] [Google Scholar]
  • 288.Regus-Leidig H, Atorf J, Feigenspan A, Kremers J, Maw MA, Brandstätter JH. Photoreceptor degeneration in two mouse models for congenital stationary night blindness type 2. PLoS One 9: e86769, 2014. doi: 10.1371/journal.pone.0086769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Regus-Leidig H, Brandstätter JH. Structure and function of a complex sensory synapse. Acta Physiol (Oxf) 204: 479–486, 2012. doi: 10.1111/j.1748-1716.2011.02355.x. [DOI] [PubMed] [Google Scholar]
  • 290.Regus-Leidig H, tom Dieck S, Brandstätter JH. Absence of functional active zone protein Bassoon affects assembly and transport of ribbon precursors during early steps of photoreceptor synaptogenesis. Eur J Cell Biol 89: 468–475, 2010. doi: 10.1016/j.ejcb.2009.12.006. [DOI] [PubMed] [Google Scholar]
  • 291.Regus-Leidig H, Fuchs M, Löhner M, Leist SR, Leal-Ortiz S, Chiodo VA, Hauswirth WW, Garner CC, Brandstätter JH. In vivo knockdown of Piccolino disrupts presynaptic ribbon morphology in mouse photoreceptor synapses. Front Cell Neurosci 8: 259, 2014. doi: 10.3389/fncel.2014.00259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Regus-Leidig H, Ott C, Löhner M, Atorf J, Fuchs M, Sedmak T, Kremers J, Fejtová A, Gundelfinger ED, Brandstätter JH. Identification and immunocytochemical characterization of Piccolino, a novel Piccolo splice variant selectively expressed at sensory ribbon synapses of the eye and ear. PLoS One 8: e70373, 2013. doi: 10.1371/journal.pone.0070373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Rehman AU, Gul K, Morell RJ, Lee K, Ahmed ZM, Riazuddin S, Ali RA, Shahzad M, Jaleel A-U, Andrade PB, Khan SN, Khan S, Brewer CC, Ahmad W, Leal SM, Riazuddin S, Friedman TB. Mutations of GIPC3 cause nonsyndromic hearing loss DFNB72 but not DFNB81 that also maps to chromosome 19p. Hum Genet 130: 759–765, 2011. doi: 10.1007/s00439-011-1018-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Reim K, Regus-Leidig H, Ammermüller J, El-Kordi A, Radyushkin K, Ehrenreich H, Brandstätter JH, Brose N. Aberrant function and structure of retinal ribbon synapses in the absence of complexin 3 and complexin 4. J Cell Sci 122: 1352–1361, 2009. doi: 10.1242/jcs.045401. [DOI] [PubMed] [Google Scholar]
  • 295.Reiners J, Reidel B, El-Amraoui A, Boëda B, Huber I, Petit C, Wolfrum U. Differential distribution of harmonin isoforms and their possible role in Usher-1 protein complexes in mammalian photoreceptor cells. Invest Ophthalmol Vis Sci 44: 5006–5015, 2003. doi: 10.1167/iovs.03-0483. [DOI] [PubMed] [Google Scholar]
  • 296.Remé CE, Young RW. The effects of hibernation on cone visual cells in the ground squirrel. Invest Ophthalmol Vis Sci 16: 815–840, 1977. [PubMed] [Google Scholar]
  • 297.Rettig J, Sheng ZH, Kim DK, Hodson CD, Snutch TP, Catterall WA. Isoform-specific interaction of the alpha1A subunits of brain Ca2+ channels with the presynaptic proteins syntaxin and SNAP-25. Proc Natl Acad Sci USA 93: 7363–7368, 1996. doi: 10.1073/pnas.93.14.7363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Richards MW, Leroy J, Pratt WS, Dolphin AC. The HOOK-domain between the SH3 and the GK domains of Cavbeta subunits contains key determinants controlling calcium channel inactivation. Channels (Austin) 1: 92–101, 2007. doi: 10.4161/chan.4145. [DOI] [PubMed] [Google Scholar]
  • 299.Rieke F, Lee A, Haeseleer F. Characterization of Ca2+-binding protein 5 knockout mouse retina. Invest Ophthalmol Vis Sci 49: 5126–5135, 2008. doi: 10.1167/iovs.08-2236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Rieke F, Rudd ME. The challenges natural images pose for visual adaptation. Neuron 64: 605–616, 2009. doi: 10.1016/j.neuron.2009.11.028. [DOI] [PubMed] [Google Scholar]
  • 301.Rieke F, Schwartz EA. Asynchronous transmitter release: control of exocytosis and endocytosis at the salamander rod synapse. J Physiol 493: 1–8, 1996. doi: 10.1113/jphysiol.1996.sp021360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Roberts WM, Jacobs RA, Hudspeth AJ. Colocalization of ion channels involved in frequency selectivity and synaptic transmission at presynaptic active zones of hair cells. J Neurosci 10: 3664–3684, 1990. doi: 10.1523/JNEUROSCI.10-11-03664.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Rodriguez-Contreras A, Nonner W, Yamoah EN. Ca2+ transport properties and determinants of anomalous mole fraction effects of single voltage-gated Ca2+ channels in hair cells from bullfrog saccule. J Physiol 538: 729–745, 2002. doi: 10.1113/jphysiol.2001.013312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Rodriguez-Contreras A, Yamoah EN. Direct measurement of single-channel Ca(2+) currents in bullfrog hair cells reveals two distinct channel subtypes. J Physiol 534: 669–689, 2001. doi: 10.1111/j.1469-7793.2001.00669.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Rodríguez-Contreras A, Yamoah EN. Effects of permeant ion concentrations on the gating of L-type Ca2+ channels in hair cells. Biophys J 84: 3457–3469, 2003. doi: 10.1016/S0006-3495(03)70066-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Roux I, Safieddine S, Nouvian R, Grati M, Simmler M-C, Bahloul A, Perfettini I, Le Gall M, Rostaing P, Hamard G, Triller A, Avan P, Moser T, Petit C. Otoferlin, defective in a human deafness form, is essential for exocytosis at the auditory ribbon synapse. Cell 127: 277–289, 2006. doi: 10.1016/j.cell.2006.08.040. [DOI] [PubMed] [Google Scholar]
  • 308.Rutherford MA, Chapochnikov NM, Moser T. Spike encoding of neurotransmitter release timing by spiral ganglion neurons of the cochlea. J Neurosci 32: 4773–4789, 2012. doi: 10.1523/JNEUROSCI.4511-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Rutherford MA, Roberts WM. Frequency selectivity of synaptic exocytosis in frog saccular hair cells. Proc Natl Acad Sci USA 103: 2898–2903, 2006. doi: 10.1073/pnas.0511005103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Sampath AP, Rieke F. Selective transmission of single photon responses by saturation at the rod-to-rod bipolar synapse. Neuron 41: 431–443, 2004. doi: 10.1016/S0896-6273(04)00005-4. [DOI] [PubMed] [Google Scholar]
  • 311.Sang L, Dick IE, Yue DT. Protein kinase A modulation of CaV1.4 calcium channels. Nat Commun 7: 12239, 2016. doi: 10.1038/ncomms12239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Scharinger A, Eckrich S, Vandael DH, Schönig K, Koschak A, Hecker D, Kaur G, Lee A, Sah A, Bartsch D, Benedetti B, Lieb A, Schick B, Singewald N, Sinnegger-Brauns MJ, Carbone E, Engel J, Striessnig J. Cell-type-specific tuning of Cav1.3 Ca(2+)-channels by a C-terminal automodulatory domain. Front Cell Neurosci 9: 309, 2015. doi: 10.3389/fncel.2015.00309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313.Schmitz F, Königstorfer A, Südhof TC. RIBEYE, a component of synaptic ribbons: a protein’s journey through evolution provides insight into synaptic ribbon function. Neuron 28: 857–872, 2000. doi: 10.1016/S0896-6273(00)00159-8. [DOI] [PubMed] [Google Scholar]
  • 314.Schnee ME, Ricci AJ. Biophysical and pharmacological characterization of voltage-gated calcium currents in turtle auditory hair cells. J Physiol 549: 697–717, 2003. doi: 10.1113/jphysiol.2002.037481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Schnee ME, Santos-Sacchi J, Castellano-Muñoz M, Kong J-H, Ricci AJ. Calcium-dependent synaptic vesicle trafficking underlies indefatigable release at the hair cell afferent fiber synapse. Neuron 70: 326–338, 2011. doi: 10.1016/j.neuron.2011.01.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Schneggenburger R, Neher E. Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature 406: 889–893, 2000. doi: 10.1038/35022702. [DOI] [PubMed] [Google Scholar]
  • 317.Schneggenburger R, Neher E. Presynaptic calcium and control of vesicle fusion. Curr Opin Neurobiol 15: 266–274, 2005. doi: 10.1016/j.conb.2005.05.006. [DOI] [PubMed] [Google Scholar]
  • 318.Schrauwen I, Helfmann S, Inagaki A, Predoehl F, Tabatabaiefar MA, Picher MM, Sommen M, Zazo Seco C, Oostrik J, Kremer H, Dheedene A, Claes C, Fransen E, Chaleshtori MH, Coucke P, Lee A, Moser T, Van Camp G. A mutation in CABP2, expressed in cochlear hair cells, causes autosomal-recessive hearing impairment. Am J Hum Genet 91: 636–645, 2012. doi: 10.1016/j.ajhg.2012.08.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Schwarz K, Natarajan S, Kassas N, Vitale N, Schmitz F. The synaptic ribbon is a site of phosphatidic acid generation in ribbon synapses. J Neurosci 31: 15996–16011, 2011. doi: 10.1523/JNEUROSCI.2965-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Sendin G, Bourien J, Rassendren F, Puel J-L, Nouvian R. Spatiotemporal pattern of action potential firing in developing inner hair cells of the mouse cochlea. Proc Natl Acad Sci USA 111: 1999–2004, 2014. doi: 10.1073/pnas.1319615111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Sendin G, Bulankina AV, Riedel D, Moser T. Maturation of ribbon synapses in hair cells is driven by thyroid hormone. J Neurosci 27: 3163–3173, 2007. doi: 10.1523/JNEUROSCI.3974-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Shaltiel L, Paparizos C, Fenske S, Hassan S, Gruner C, Rötzer K, Biel M, Wahl-Schott CA. Complex regulation of voltage-dependent activation and inactivation properties of retinal voltage-gated Cav1.4 L-type Ca2+ channels by Ca2+-binding protein 4 (CaBP4). J Biol Chem 287: 36312–36321, 2012. doi: 10.1074/jbc.M112.392811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Sheets L, He XJ, Olt J, Schreck M, Petralia RS, Wang Y-X, Zhang Q, Beirl A, Nicolson T, Marcotti W, Trapani JG, Kindt KS. Enlargement of Ribbons in Zebrafish Hair Cells Increases Calcium Currents But Disrupts Afferent Spontaneous Activity and Timing of Stimulus Onset. J Neurosci 37: 6299–6313, 2017. doi: 10.1523/JNEUROSCI.2878-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Sheets L, Kindt KS, Nicolson T. Presynaptic CaV1.3 channels regulate synaptic ribbon size and are required for synaptic maintenance in sensory hair cells. J Neurosci 32: 17273–17286, 2012. doi: 10.1523/JNEUROSCI.3005-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Sheets L, Trapani JG, Mo W, Obholzer N, Nicolson T. Ribeye is required for presynaptic Ca(V)1.3a channel localization and afferent innervation of sensory hair cells. Development 138: 1309–1319, 2011. doi: 10.1242/dev.059451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Shen Y, Yu D, Hiel H, Liao P, Yue DT, Fuchs PA, Soong TW. Alternative splicing of the Ca(v)1.3 channel IQ domain, a molecular switch for Ca2+-dependent inactivation within auditory hair cells. J Neurosci 26: 10690–10699, 2006. doi: 10.1523/JNEUROSCI.2093-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Sheng ZH, Rettig J, Cook T, Catterall WA. Calcium-dependent interaction of N-type calcium channels with the synaptic core complex. Nature 379: 451–454, 1996. doi: 10.1038/379451a0. [DOI] [PubMed] [Google Scholar]
  • 328.Sheng ZH, Rettig J, Takahashi M, Catterall WA. Identification of a syntaxin-binding site on N-type calcium channels. Neuron 13: 1303–1313, 1994. doi: 10.1016/0896-6273(94)90417-0. [DOI] [PubMed] [Google Scholar]
  • 329.Simms BA, Souza IA, Rehak R, Zamponi GW. The Cav1.2 N terminus contains a CaM kinase site that modulates channel trafficking and function. Pflugers Arch 467: 677–686, 2015. doi: 10.1007/s00424-014-1538-7. [DOI] [PubMed] [Google Scholar]
  • 330.Simms BA, Zamponi GW. Trafficking and stability of voltage-gated calcium channels. Cell Mol Life Sci 69: 843–856, 2012. doi: 10.1007/s00018-011-0843-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Singer JH, Lassová L, Vardi N, Diamond JS. Coordinated multivesicular release at a mammalian ribbon synapse. Nat Neurosci 7: 826–833, 2004. doi: 10.1038/nn1280. [DOI] [PubMed] [Google Scholar]
  • 332.Singh A, Gebhart M, Fritsch R, Sinnegger-Brauns MJ, Poggiani C, Hoda J-C, Engel J, Romanin C, Striessnig J, Koschak A. Modulation of voltage- and Ca2+-dependent gating of CaV1.3 L-type calcium channels by alternative splicing of a C-terminal regulatory domain. J Biol Chem 283: 20733–20744, 2008. doi: 10.1074/jbc.M802254200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Singh A, Hamedinger D, Hoda J-C, Gebhart M, Koschak A, Romanin C, Striessnig J. C-terminal modulator controls Ca2+-dependent gating of Ca(v)1.4 L-type Ca2+ channels. Nat Neurosci 9: 1108–1116, 2006. doi: 10.1038/nn1751. [DOI] [PubMed] [Google Scholar]
  • 334.Sinha R, Lee A, Rieke F, Haeseleer F. Lack of CaBP1/Caldendrin or CaBP2 Leads to Altered Ganglion Cell Responses. eNeuro 3: 0099-16.2016, 2016. doi: 10.1523/ENEURO.0099-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Snellman J, Mehta B, Babai N, Bartoletti TM, Akmentin W, Francis A, Matthews G, Thoreson W, Zenisek D. Acute destruction of the synaptic ribbon reveals a role for the ribbon in vesicle priming. Nat Neurosci 14: 1135–1141, 2011. doi: 10.1038/nn.2870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Sokal I, Haeseleer F. Insight into the role of Ca2+-binding protein 5 in vesicle exocytosis. Invest Ophthalmol Vis Sci 52: 9131–9141, 2011. doi: 10.1167/iovs.11-8246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Specht D, Tom Dieck S, Ammermüller J, Regus-Leidig H, Gundelfinger ED, Brandstätter JH. Structural and functional remodeling in the retina of a mouse with a photoreceptor synaptopathy: plasticity in the rod and degeneration in the cone system. Eur J Neurosci 26: 2506–2515, 2007. doi: 10.1111/j.1460-9568.2007.05886.x. [DOI] [PubMed] [Google Scholar]
  • 338.Specht D, Wu S-B, Turner P, Dearden P, Koentgen F, Wolfrum U, Maw M, Brandstätter JH, tom Dieck S. Effects of presynaptic mutations on a postsynaptic Cacna1s calcium channel colocalized with mGluR6 at mouse photoreceptor ribbon synapses. Invest Ophthalmol Vis Sci 50: 505–515, 2009. doi: 10.1167/iovs.08-2758. [DOI] [PubMed] [Google Scholar]
  • 339.Stella SL Jr, Thoreson WB. Differential modulation of rod and cone calcium currents in tiger salamander retina by D2 dopamine receptors and cAMP. Eur J Neurosci 12: 3537–3548, 2000. doi: 10.1046/j.1460-9568.2000.00235.x. [DOI] [PubMed] [Google Scholar]
  • 340.Sterling P, Freed MA, Smith RG. Architecture of rod and cone circuits to the on-beta ganglion cell. J Neurosci 8: 623–642, 1988. doi: 10.1523/JNEUROSCI.08-02-00623.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Sterling P, Matthews G. Structure and function of ribbon synapses. Trends Neurosci 28: 20–29, 2005. doi: 10.1016/j.tins.2004.11.009. [DOI] [PubMed] [Google Scholar]
  • 342.Stockner T, Koschak A. What can naturally occurring mutations tell us about Ca(v)1.x channel function? Biochim Biophys Acta 1828: 1598–1607, 2013. doi: 10.1016/j.bbamem.2012.11.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343.Straiker A, Sullivan JM. Cannabinoid receptor activation differentially modulates ion channels in photoreceptors of the tiger salamander. J Neurophysiol 89: 2647–2654, 2003. doi: 10.1152/jn.00268.2002. [DOI] [PubMed] [Google Scholar]
  • 344.Strenzke N, Chakrabarti R, Al-Moyed H, Müller A, Hoch G, Pangrsic T, Yamanbaeva G, Lenz C, Pan K-T, Auge E, Geiss-Friedlander R, Urlaub H, Brose N, Wichmann C, Reisinger E. Hair cell synaptic dysfunction, auditory fatigue and thermal sensitivity in otoferlin Ile515Thr mutants. EMBO J 35: 2519–2535, 2016. doi: 10.15252/embj.201694564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 345.Strenzke N, Chanda S, Kopp-Scheinpflug C, Khimich D, Reim K, Bulankina AV, Neef A, Wolf F, Brose N, Xu-Friedman MA, Moser T. Complexin-I is required for high-fidelity transmission at the endbulb of Held auditory synapse. J Neurosci 29: 7991–8004, 2009. doi: 10.1523/JNEUROSCI.0632-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.Strettoi E, Pignatelli V. Modifications of retinal neurons in a mouse model of retinitis pigmentosa. Proc Natl Acad Sci USA 97: 11020–11025, 2000. doi: 10.1073/pnas.190291097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.Strettoi E, Pignatelli V, Rossi C, Porciatti V, Falsini B. Remodeling of second-order neurons in the retina of rd/rd mutant mice. Vision Res 43: 867–877, 2003. doi: 10.1016/S0042-6989(02)00594-1. [DOI] [PubMed] [Google Scholar]
  • 348.Strettoi E, Porciatti V, Falsini B, Pignatelli V, Rossi C. Morphological and functional abnormalities in the inner retina of the rd/rd mouse. J Neurosci 22: 5492–5504, 2002. doi: 10.1523/JNEUROSCI.22-13-05492.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 349.Strom TM, Nyakatura G, Apfelstedt-Sylla E, Hellebrand H, Lorenz B, Weber BH, Wutz K, Gutwillinger N, Rüther K, Drescher B, Sauer C, Zrenner E, Meitinger T, Rosenthal A, Meindl A. An L-type calcium-channel gene mutated in incomplete X-linked congenital stationary night blindness. Nat Genet 19: 260–263, 1998. doi: 10.1038/940. [DOI] [PubMed] [Google Scholar]
  • 350.Su ZL, Jiang SC, Gu R, Yang WP. Two types of calcium channels in bullfrog saccular hair cells. Hear Res 87: 62–68, 1995. doi: 10.1016/0378-5955(95)00079-J. [DOI] [PubMed] [Google Scholar]
  • 351.Südhof TC. The presynaptic active zone. Neuron 75: 11–25, 2012. doi: 10.1016/j.neuron.2012.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 352.Sugawara Y. Two Ca current components of the receptor current in the electroreceptors of the marine catfish Plotosus. J Gen Physiol 93: 365–380, 1989. doi: 10.1085/jgp.93.2.365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353.Szmajda BA, Devries SH. Glutamate spillover between mammalian cone photoreceptors. J Neurosci 31: 13431–13441, 2011. doi: 10.1523/JNEUROSCI.2105-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 354.Tabatabaiefar MA, Alasti F, Shariati L, Farrokhi E, Fransen E, Nooridaloii MR, Chaleshtori MH, Van Camp G. DFNB93, a novel locus for autosomal recessive moderate-to-severe hearing impairment. Clin Genet 79: 594–598, 2011. doi: 10.1111/j.1399-0004.2010.01593.x. [DOI] [PubMed] [Google Scholar]
  • 355.Taberner AM, Liberman MC. Response properties of single auditory nerve fibers in the mouse. J Neurophysiol 93: 557–569, 2005. doi: 10.1152/jn.00574.2004. [DOI] [PubMed] [Google Scholar]
  • 356.Tadross MR, Dick IE, Yue DT. Mechanism of local and global Ca2+ sensing by calmodulin in complex with a Ca2+ channel. Cell 133: 1228–1240, 2008. doi: 10.1016/j.cell.2008.05.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 357.Taiakina V, Boone AN, Fux J, Senatore A, Weber-Adrian D, Guillemette JG, Spafford JD. The calmodulin-binding, short linear motif, NSCaTE is conserved in L-type channel ancestors of vertebrate Cav1.2 and Cav1.3 channels. PLoS One 8: e61765, 2013. doi: 10.1371/journal.pone.0061765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Tan BZ, Jiang F, Tan MY, Yu D, Huang H, Shen Y, Soong TW. Functional characterization of alternative splicing in the C terminus of L-type CaV1.3 channels. J Biol Chem 286: 42725–42735, 2011. doi: 10.1074/jbc.M111.265207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.Tan GMY, Yu D, Wang J, Soong TW. Alternative splicing at C terminus of Ca(V)1.4 calcium channel modulates calcium-dependent inactivation, activation potential, and current density. J Biol Chem 287: 832–847, 2012. doi: 10.1074/jbc.M111.268722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Taylor WR, Morgans C. Localization and properties of voltage-gated calcium channels in cone photoreceptors of Tupaia belangeri. Vis Neurosci 15: 541–552, 1998. doi: 10.1017/S0952523898153142. [DOI] [PubMed] [Google Scholar]
  • 361.Thoreson WB, Babai N, Bartoletti TM. Feedback from horizontal cells to rod photoreceptors in vertebrate retina. J Neurosci 28: 5691–5695, 2008. doi: 10.1523/JNEUROSCI.0403-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Thoreson WB, Bryson EJ, Rabl K. Reciprocal interactions between calcium and chloride in rod photoreceptors. J Neurophysiol 90: 1747–1753, 2003. doi: 10.1152/jn.00932.2002. [DOI] [PubMed] [Google Scholar]
  • 363.Thoreson WB, Mercer AJ, Cork KM, Szalewski RJ. Lateral mobility of L-type calcium channels in synaptic terminals of retinal bipolar cells. Mol Vis 19: 16–24, 2013. [PMC free article] [PubMed] [Google Scholar]
  • 364.Thoreson WB, Miller RF. Removal of extracellular chloride suppresses transmitter release from photoreceptor terminals in the mudpuppy retina. J Gen Physiol 107: 631–642, 1996. doi: 10.1085/jgp.107.5.631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Thoreson WB, Nitzan R, Miller RF. Reducing extracellular Cl suppresses dihydropyridine-sensitive Ca2+ currents and synaptic transmission in amphibian photoreceptors. J Neurophysiol 77: 2175–2190, 1997. doi: 10.1152/jn.1997.77.4.2175. [DOI] [PubMed] [Google Scholar]
  • 366.Thoreson WB, Nitzan R, Miller RF. Chloride efflux inhibits single calcium channel open probability in vertebrate photoreceptors: chloride imaging and cell-attached patch-clamp recordings. Vis Neurosci 17: 197–206, 2000. doi: 10.1017/S0952523800172025. [DOI] [PubMed] [Google Scholar]
  • 367.Thoreson WB, Rabl K, Townes-Anderson E, Heidelberger R. A highly Ca2+-sensitive pool of vesicles contributes to linearity at the rod photoreceptor ribbon synapse. Neuron 42: 595–605, 2004. doi: 10.1016/S0896-6273(04)00254-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367a.tom Dieck S, Altrock WD, Kessels MM, Qualmann B, Regus H, Brauner D, Fejtová A, Bracko O, Gundelfinger ED, Brandstätter JH. Molecular dissection of the photoreceptor ribbon synapse: physical interaction of Bassoon and RIBEYE is essential for the assembly of the ribbon complex. J Cell Biol 168: 825–836, 2005. doi: 10.1083/jcb.200408157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367b.tom Dieck S, Brandstätter JH. Ribbon synapses of the retina. Cell Tissue Res 326: 339–346, 2006. doi: 10.1007/s00441-006-0234-0. [DOI] [PubMed] [Google Scholar]
  • 367c.tom Dieck S, Specht D, Strenzke N, Hida Y, Krishnamoorthy V, Schmidt K-F, Inoue E, Ishizaki H, Tanaka-Okamoto M, Miyoshi J, Hagiwara A, Brandstätter JH, Löwel S, Gollisch T, Ohtsuka T, Moser T. Deletion of the presynaptic scaffold CAST reduces active zone size in rod photoreceptors and impairs visual processing. J Neurosci 32: 12192–12203, 2012. doi: 10.1523/JNEUROSCI.0752-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 368.Trexler EB, Casti ARR, Zhang Y. Nonlinearity and noise at the rod-rod bipolar cell synapse. Vis Neurosci 28: 61–68, 2011. doi: 10.1017/S0952523810000301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369.Trimbuch T, Rosenmund C. Should I stop or should I go? The role of complexin in neurotransmitter release. Nat Rev Neurosci 17: 118–125, 2016. doi: 10.1038/nrn.2015.16. [DOI] [PubMed] [Google Scholar]
  • 370.Tritsch NX, Bergles DE. Developmental regulation of spontaneous activity in the Mammalian cochlea. J Neurosci 30: 1539–1550, 2010. doi: 10.1523/JNEUROSCI.3875-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 371.Tritsch NX, Rodríguez-Contreras A, Crins TTH, Wang HC, Borst JGG, Bergles DE. Calcium action potentials in hair cells pattern auditory neuron activity before hearing onset. Nat Neurosci 13: 1050–1052, 2010. doi: 10.1038/nn.2604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 372.Tritsch NX, Yi E, Gale JE, Glowatzki E, Bergles DE. The origin of spontaneous activity in the developing auditory system. Nature 450: 50–55, 2007. doi: 10.1038/nature06233. [DOI] [PubMed] [Google Scholar]
  • 373.Tsukamoto Y, Morigiwa K, Ueda M, Sterling P. Microcircuits for night vision in mouse retina. J Neurosci 21: 8616–8623, 2001. doi: 10.1523/JNEUROSCI.21-21-08616.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 374.Tucker T, Fettiplace R. Confocal imaging of calcium microdomains and calcium extrusion in turtle hair cells. Neuron 15: 1323–1335, 1995. doi: 10.1016/0896-6273(95)90011-X. [DOI] [PubMed] [Google Scholar]
  • 375.Uchida S, Yamada S, Nagai K, Deguchi Y, Kimura R. Brain pharmacokinetics and in vivo receptor binding of 1,4-dihydropyridine calcium channel antagonists. Life Sci 61: 2083–2090, 1997. doi: 10.1016/S0024-3205(97)00881-3. [DOI] [PubMed] [Google Scholar]
  • 376.Ullrich B, Südhof TC. Distribution of synaptic markers in the retina: implications for synaptic vesicle traffic in ribbon synapses. J Physiol Paris 88: 249–257, 1994. doi: 10.1016/0928-4257(94)90088-4. [DOI] [PubMed] [Google Scholar]
  • 377.Uthaiah RC, Hudspeth AJ. Molecular anatomy of the hair cell’s ribbon synapse. J Neurosci 30: 12387–12399, 2010. doi: 10.1523/JNEUROSCI.1014-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Vaithianathan T, Henry D, Akmentin W, Matthews G. Functional roles of complexin in neurotransmitter release at ribbon synapses of mouse retinal bipolar neurons. J Neurosci 35: 4065–4070, 2015. doi: 10.1523/JNEUROSCI.2703-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.Van Hook MJ, Parmelee CM, Chen M, Cork KM, Curto C, Thoreson WB. Calmodulin enhances ribbon replenishment and shapes filtering of synaptic transmission by cone photoreceptors. J Gen Physiol 144: 357–378, 2014. doi: 10.1085/jgp.201411229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379a.Van Rossum MC, Smith RG. Noise removal at the rod synapse of mammalian retina. Vis Neurosci 15: 809–821, 1998. doi: 10.1017/S0952523898155037. [DOI] [PubMed] [Google Scholar]
  • 380.Vellani V, Reynolds AM, McNaughton PA. Modulation of the synaptic Ca2+ current in salamander photoreceptors by polyunsaturated fatty acids and retinoids. J Physiol 529: 333–344, 2000. doi: 10.1111/j.1469-7793.2000.00333.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Vincent PFY, Bouleau Y, Charpentier G, Emptoz A, Safieddine S, Petit C, Dulon D. Different CaV1.3 Channel Isoforms Control Distinct Components of the Synaptic Vesicle Cycle in Auditory Inner Hair Cells. J Neurosci 37: 2960–2975, 2017. doi: 10.1523/JNEUROSCI.2374-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Vincent PFY, Bouleau Y, Safieddine S, Petit C, Dulon D. Exocytotic machineries of vestibular type I and cochlear ribbon synapses display similar intrinsic otoferlin-dependent Ca2+ sensitivity but a different coupling to Ca2+ channels. J Neurosci 34: 10853–10869, 2014. doi: 10.1523/JNEUROSCI.0947-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Vogl C, Cooper BH, Neef J, Wojcik SM, Reim K, Reisinger E, Brose N, Rhee J-S, Moser T, Wichmann C. Unconventional molecular regulation of synaptic vesicle replenishment in cochlear inner hair cells. J Cell Sci 128: 638–644, 2015. doi: 10.1242/jcs.162099. [DOI] [PubMed] [Google Scholar]
  • 384.Vogl C, Panou I, Yamanbaeva G, Wichmann C, Mangosing SJ, Vilardi F, Indzhykulian AA, Pangršič T, Santarelli R, Rodriguez-Ballesteros M, Weber T, Jung S, Cardenas E, Wu X, Wojcik SM, Kwan KY, Del Castillo I, Schwappach B, Strenzke N, Corey DP, Lin S-Y, Moser T. Tryptophan-rich basic protein (WRB) mediates insertion of the tail-anchored protein otoferlin and is required for hair cell exocytosis and hearing. EMBO J 35: 2536–2552, 2016. doi: 10.15252/embj.201593565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Vollrath L, Spiwoks-Becker I. Plasticity of retinal ribbon synapses. Microsc Res Tech 35: 472–487, 1996. doi:. [DOI] [PubMed] [Google Scholar]
  • 386.Wahl-Schott C, Baumann L, Cuny H, Eckert C, Griessmeier K, Biel M. Switching off calcium-dependent inactivation in L-type calcium channels by an autoinhibitory domain. Proc Natl Acad Sci USA 103: 15657–15662, 2006. doi: 10.1073/pnas.0604621103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 387.Waites CL, Leal-Ortiz SA, Okerlund N, Dalke H, Fejtova A, Altrock WD, Gundelfinger ED, Garner CC. Bassoon and Piccolo maintain synapse integrity by regulating protein ubiquitination and degradation. EMBO J 32: 954–969, 2013. doi: 10.1038/emboj.2013.27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 388.Waithe D, Ferron L, Page KM, Chaggar K, Dolphin AC. Beta-subunits promote the expression of Ca(V)2.2 channels by reducing their proteasomal degradation. J Biol Chem 286: 9598–9611, 2011. doi: 10.1074/jbc.M110.195909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Wang HC, Lin C-C, Cheung R, Zhang-Hooks Y, Agarwal A, Ellis-Davies G, Rock J, Bergles DE. Spontaneous Activity of Cochlear Hair Cells Triggered by Fluid Secretion Mechanism in Adjacent Support Cells. Cell 163: 1348–1359, 2015. doi: 10.1016/j.cell.2015.10.070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 390.Wang H-G, George MS, Kim J, Wang C, Pitt GS. Ca2+/calmodulin regulates trafficking of Ca(V)1.2 Ca2+ channels in cultured hippocampal neurons. J Neurosci 27: 9086–9093, 2007. doi: 10.1523/JNEUROSCI.1720-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 391.Wang SSH, Held RG, Wong MY, Liu C, Karakhanyan A, Kaeser PS. Fusion Competent Synaptic Vesicles Persist upon Active Zone Disruption and Loss of Vesicle Docking. Neuron 91: 777–791, 2016. doi: 10.1016/j.neuron.2016.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 392.Wang T-M, Holzhausen LC, Kramer RH. Imaging an optogenetic pH sensor reveals that protons mediate lateral inhibition in the retina. Nat Neurosci 17: 262–268, 2014. doi: 10.1038/nn.3627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 393.Wang Y, Fehlhaber KE, Sarria I, Cao Y, Ingram NT, Guerrero-Given D, Throesch B, Baldwin K, Kamasawa N, Ohtsuka T, Sampath AP, Martemyanov KA. The Auxiliary Calcium Channel Subunit α2δ4 Is Required for Axonal Elaboration, Synaptic Transmission, and Wiring of Rod Photoreceptors. Neuron 93: 1359–1374.e6, 2017. doi: 10.1016/j.neuron.2017.02.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 394.Wang Y, Okamoto M, Schmitz F, Hofmann K, Südhof TC. Rim is a putative Rab3 effector in regulating synaptic-vesicle fusion. Nature 388: 593–598, 1997. doi: 10.1038/41580. [DOI] [PubMed] [Google Scholar]
  • 395.Wang Y, Sugita S, Südhof TC. The RIM/NIM family of neuronal C2 domain proteins. Interactions with Rab3 and a new class of Src homology 3 domain proteins. J Biol Chem 275: 20033–20044, 2000. doi: 10.1074/jbc.M909008199. [DOI] [PubMed] [Google Scholar]
  • 396.Warren TJ, Van Hook MJ, Supuran CT, Thoreson WB. Sources of protons and a role for bicarbonate in inhibitory feedback from horizontal cells to cones in Ambystoma tigrinum retina. J Physiol 594: 6661–6677, 2016. doi: 10.1113/JP272533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 397.Weiler S, Krinner S, Wong AB, Moser T, Pangršič T. ATP hydrolysis is critically required for function of CaV1.3 channels in cochlear inner hair cells via fueling Ca2+ clearance. J Neurosci 34: 6843–6848, 2014. doi: 10.1523/JNEUROSCI.4990-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 398.Wilkinson MF, Barnes S. The dihydropyridine-sensitive calcium channel subtype in cone photoreceptors. J Gen Physiol 107: 621–630, 1996. doi: 10.1085/jgp.107.5.621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 399.Wingard JN, Chan J, Bosanac I, Haeseleer F, Palczewski K, Ikura M, Ames JB. Structural analysis of Mg2+ and Ca2+ binding to CaBP1, a neuron-specific regulator of calcium channels. J Biol Chem 280: 37461–37470, 2005. doi: 10.1074/jbc.M508541200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 400.Winter IM, Robertson D, Yates GK. Diversity of characteristic frequency rate-intensity functions in guinea pig auditory nerve fibres. Hear Res 45: 191–202, 1990. doi: 10.1016/0378-5955(90)90120-E. [DOI] [PubMed] [Google Scholar]
  • 401.Wong AB, Jing Z, Rutherford MA, Frank T, Strenzke N, Moser T. Concurrent maturation of inner hair cell synaptic Ca2+ influx and auditory nerve spontaneous activity around hearing onset in mice. J Neurosci 33: 10661–10666, 2013. doi: 10.1523/JNEUROSCI.1215-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 402.Wong AB, Rutherford MA, Gabrielaitis M, Pangršič T, Göttfert F, Frank T, Michanski S, Hell S, Wolf F, Wichmann C, Moser T. Developmental refinement of hair cell synapses tightens the coupling of Ca2+ influx to exocytosis. EMBO J 33: 247–264, 2014. doi: 10.1002/embj.201387110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.Wu SM, Qiao X, Noebels JL, Yang XL. Localization and modulatory actions of zinc in vertebrate retina. Vision Res 33: 2611–2616, 1993. doi: 10.1016/0042-6989(93)90219-M. [DOI] [PubMed] [Google Scholar]
  • 404.Wycisk KA, Budde B, Feil S, Skosyrski S, Buzzi F, Neidhardt J, Glaus E, Nürnberg P, Ruether K, Berger W. Structural and functional abnormalities of retinal ribbon synapses due to Cacna2d4 mutation. Invest Ophthalmol Vis Sci 47: 3523–3530, 2006. doi: 10.1167/iovs.06-0271. [DOI] [PubMed] [Google Scholar]
  • 405.Wycisk KA, Zeitz C, Feil S, Wittmer M, Forster U, Neidhardt J, Wissinger B, Zrenner E, Wilke R, Kohl S, Berger W. Mutation in the auxiliary calcium-channel subunit CACNA2D4 causes autosomal recessive cone dystrophy. Am J Hum Genet 79: 973–977, 2006. doi: 10.1086/508944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Xiao H, Chen X, Steele EC Jr. Abundant L-type calcium channel Ca(v)1.3 (alpha1D) subunit mRNA is detected in rod photoreceptors of the mouse retina via in situ hybridization. Mol Vis 13: 764–771, 2007. [PMC free article] [PubMed] [Google Scholar]
  • 407.Xu JW, Slaughter MM. Large-conductance calcium-activated potassium channels facilitate transmitter release in salamander rod synapse. J Neurosci 25: 7660–7668, 2005. doi: 10.1523/JNEUROSCI.1572-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 408.Xu W, Lipscombe D. Neuronal Ca(V)1.3α(1) L-type channels activate at relatively hyperpolarized membrane potentials and are incompletely inhibited by dihydropyridines. J Neurosci 21: 5944–5951, 2001. doi: 10.1523/JNEUROSCI.21-16-05944.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Yang PS, Alseikhan BA, Hiel H, Grant L, Mori MX, Yang W, Fuchs PA, Yue DT. Switching of Ca2+-dependent inactivation of Ca(v)1.3 channels by calcium binding proteins of auditory hair cells. J Neurosci 26: 10677–10689, 2006. doi: 10.1523/JNEUROSCI.3236-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 410.Yang PS, Johny MB, Yue DT. Allostery in Ca2+ channel modulation by calcium-binding proteins. Nat Chem Biol 10: 231–238, 2014. doi: 10.1038/nchembio.1436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 411.Yang T, Scholl ES, Pan N, Fritzsch B, Haeseleer F, Lee A. Expression and Localization of CaBP Ca2+ Binding Proteins in the Mouse Cochlea. PLoS One 11: e0147495, 2016. doi: 10.1371/journal.pone.0147495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 412.Yang Y-M, Fedchyshyn MJ, Grande G, Aitoubah J, Tsang CW, Xie H, Ackerley CA, Trimble WS, Wang L-Y. Septins regulate developmental switching from microdomain to nanodomain coupling of Ca(2+) influx to neurotransmitter release at a central synapse. Neuron 67: 100–115, 2010. doi: 10.1016/j.neuron.2010.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 413.Yates GK, Winter IM, Robertson D. Basilar membrane nonlinearity determines auditory nerve rate-intensity functions and cochlear dynamic range. Hear Res 45: 203–219, 1990. doi: 10.1016/0378-5955(90)90121-5. [DOI] [PubMed] [Google Scholar]
  • 414.Zabouri N, Haverkamp S. Calcium channel-dependent molecular maturation of photoreceptor synapses. PLoS One 8: e63853, 2013. doi: 10.1371/journal.pone.0063853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 415.Zagaeski M, Cody AR, Russell IJ, Mountain DC. Transfer characteristic of the inner hair cell synapse: steady-state analysis. J Acoust Soc Am 95: 3430–3434, 1994. doi: 10.1121/1.409963. [DOI] [PubMed] [Google Scholar]
  • 416.Zampini V, Johnson SL, Franz C, Knipper M, Holley MC, Magistretti J, Masetto S, Marcotti W. Burst activity and ultrafast activation kinetics of CaV1.3 Ca2+ channels support presynaptic activity in adult gerbil hair cell ribbon synapses. J Physiol 591: 3811–3820, 2013. doi: 10.1113/jphysiol.2013.251272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 417.Zampini V, Johnson SL, Franz C, Knipper M, Holley MC, Magistretti J, Russo G, Marcotti W, Masetto S. Fine tuning of CaV1.3 Ca2+ channel properties in adult inner hair cells positioned in the most sensitive region of the gerbil cochlea. PLoS One 9: e113750, 2014. doi: 10.1371/journal.pone.0113750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 418.Zampini V, Johnson SL, Franz C, Lawrence ND, Münkner S, Engel J, Knipper M, Magistretti J, Masetto S, Marcotti W. Elementary properties of CaV1.3 Ca(2+) channels expressed in mouse cochlear inner hair cells. J Physiol 588: 187–199, 2010. doi: 10.1113/jphysiol.2009.181917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 419.Zamponi GW, Striessnig J, Koschak A, Dolphin AC. The Physiology, Pathology, and Pharmacology of Voltage-Gated Calcium Channels and Their Future Therapeutic Potential. Pharmacol Rev 67: 821–870, 2015. doi: 10.1124/pr.114.009654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 420.Zeitz C, Kloeckener-Gruissem B, Forster U, Kohl S, Magyar I, Wissinger B, Mátyás G, Borruat F-X, Schorderet DF, Zrenner E, Munier FL, Berger W. Mutations in CABP4, the gene encoding the Ca2+-binding protein 4, cause autosomal recessive night blindness. Am J Hum Genet 79: 657–667, 2006. doi: 10.1086/508067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 421.Zeitz C, Robson AG, Audo I. Congenital stationary night blindness: an analysis and update of genotype-phenotype correlations and pathogenic mechanisms. Prog Retin Eye Res 45: 58–110, 2015. doi: 10.1016/j.preteyeres.2014.09.001. [DOI] [PubMed] [Google Scholar]
  • 422.Zenisek D, Horst NK, Merrifield C, Sterling P, Matthews G. Visualizing synaptic ribbons in the living cell. J Neurosci 24: 9752–9759, 2004. doi: 10.1523/JNEUROSCI.2886-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 423.Zhou H, Yu K, McCoy KL, Lee A. Molecular mechanism for divergent regulation of Cav1.2 Ca2+ channels by calmodulin and Ca2+-binding protein-1. J Biol Chem 280: 29612–29619, 2005. doi: 10.1074/jbc.M504167200. [DOI] [PubMed] [Google Scholar]
  • 424.Zidanic M, Fuchs PA. Kinetic analysis of barium currents in chick cochlear hair cells. Biophys J 68: 1323–1336, 1995. doi: 10.1016/S0006-3495(95)80305-X. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Physiological Reviews are provided here courtesy of American Physiological Society

RESOURCES