Skip to main content

REVIEW article

Front. Cell. Neurosci., 21 September 2017
Sec. Cellular Neurophysiology
Volume 11 - 2017 | https://doi.org/10.3389/fncel.2017.00297

Zinc as a Neuromodulator in the Central Nervous System with a Focus on the Olfactory Bulb

  • 1Program in Neuroscience, Florida State University, Tallahassee, FL, United States
  • 2Department of Biological Science, Florida State University, Tallahassee, FL, United States

The olfactory bulb (OB) is central to the sense of smell, as it is the site of the first synaptic relay involved in the processing of odor information. Odor sensations are first transduced by olfactory sensory neurons (OSNs) before being transmitted, by way of the OB, to higher olfactory centers that mediate olfactory discrimination and perception. Zinc is a common trace element, and it is highly concentrated in the synaptic vesicles of subsets of glutamatergic neurons in some brain regions including the hippocampus and OB. In addition, zinc is contained in the synaptic vesicles of some glycinergic and GABAergic neurons. Thus, zinc released from synaptic vesicles is available to modulate synaptic transmission mediated by excitatory (e.g., N-methyl-D aspartate (NMDA), alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)) and inhibitory (e.g., gamma-aminobutyric acid (GABA), glycine) amino acid receptors. Furthermore, extracellular zinc can alter the excitability of neurons through effects on a variety of voltage-gated ion channels. Consistent with the notion that zinc acts as a regulator of neuronal activity, we and others have shown zinc modulation (inhibition and/or potentiation) of amino acid receptors and voltage-gated ion channels expressed by OB neurons. This review summarizes the locations and release of vesicular zinc in the central nervous system (CNS), including in the OB. It also summarizes the effects of zinc on various amino acid receptors and ion channels involved in regulating synaptic transmission and neuronal excitability, with a special emphasis on the actions of zinc as a neuromodulator in the OB. An understanding of how neuroactive substances such as zinc modulate receptors and ion channels expressed by OB neurons will increase our understanding of the roles that synaptic circuits in the OB play in odor information processing and transmission.

Introduction

Zinc is a common trace element that is found in multiple brain regions including the hippocampus, cerebral neocortex, hypothalamus, amygdala and olfactory bulb (OB; Donaldson et al., 1973; Gulya et al., 1991; Ono and Cherian, 1999). In addition to serving structural, catalytic, and regulatory functions, zinc is thought to play a role as a neuromodulator (Bitanihirwe and Cunningham, 2009; Nakashima and Dyck, 2009; Paoletti et al., 2009). This review summarizes the locations and release of vesicular zinc in the central nervous system (CNS), including in the OB. It also summarizes the effects of zinc on various receptors and ion channels involved in regulating neuronal excitability and synaptic transmission, with special emphasis on the actions of zinc as a neuromodulator in the OB.

Whereas most (80%–95%) zinc in the brain is tightly bound to proteins (Nakashima and Dyck, 2009; Paoletti et al., 2009), pools of chelatable zinc have been visualized with histochemical techniques and fluorescent probes (Danscher, 1981, 1996; Frederickson et al., 1983, 2000, 2005; Pérez-Clausell and Danscher, 1985; Frederickson, 1989; Blakemore et al., 2013). Although this chelatable zinc includes loosely bound zinc and ionic zinc, it primarily consists of vesicular zinc (Frederickson et al., 1983; Danscher et al., 1985; Pérez-Clausell and Danscher, 1985). A portion of zinc in the brain (15%) is contained in synaptic vesicles (Frederickson, 1989), with especially high concentrations found in the hippocampus (Danscher et al., 1985; Pérez-Clausell and Danscher, 1985; Slomianka et al., 1990) and OB (Friedman and Price, 1984; Masters et al., 1994; Jo et al., 2000; Blakemore et al., 2013). Most vesicular zinc is co-localized with glutamate in a subset of glutamatergic zinc-enriched neurons (Haug, 1967; Crawford and Connor, 1973; Frederickson et al., 1983; Pérez-Clausell and Danscher, 1985; Frederickson, 1989; Beaulieu et al., 1992). However, zinc is also contained in the synaptic vesicles of subpopulations of glycinergic and GABAergic neurons (Birinyi et al., 2001; Wang et al., 2001, 2002).

The co-localization of zinc and glutamate in synaptic vesicles, along with evidence that depolarizing stimuli release vesicular zinc (Assaf and Chung, 1984; Howell et al., 1984; Li et al., 2001; Blakemore et al., 2013), supports the view that neuronal activity can release zinc. Results from a variety of studies suggest that zinc modulates excitatory and inhibitory amino acid receptors and synaptic transmission (Smart et al., 1994; Bitanihirwe and Cunningham, 2009; Nakashima and Dyck, 2009; Paoletti et al., 2009). For example, zinc has been shown to inhibit gamma-aminobutyric acid (GABA) receptor-mediated and N-methyl-D aspartate (NMDA) receptor-mediated responses in the hippocampus (Westbrook and Mayer, 1987; Mayer and Vyklicky, 1989; Legendre and Westbrook, 1991; Xie et al., 1993; Smart et al., 1994). We have reported similar inhibition of NMDA- and GABA-receptor-mediated responses by zinc in OB neurons (Trombley and Shepherd, 1996; Trombley et al., 1998; Horning and Trombley, 2001; See Supplementary Table S1).

Various studies have explored zinc’s effects on other types of receptors and ion channels in the CNS, including metabotropic glutamate receptors, AMPA receptors (AMPARs), kainate receptors, glycine receptors, dopamine receptors, serotonin receptors, acetylcholine receptors, and P2X receptors as well as voltage-gated ion channels for Na+, K+, Ca2+, and Cl (Smart et al., 1994; Frederickson et al., 2005; Bitanihirwe and Cunningham, 2009; Nakashima and Dyck, 2009). In the OB, we (Horning and Trombley, 2001) and others (Puopolo and Belluzzi, 1998) have shown that zinc modulates several types of voltage-gated ion channels (see Supplementary Table S2). We have also demonstrated that the zinc has biphasic (potentiating/inhibiting) concentration-dependent effects on subsets of glycine receptors (Trombley and Shepherd, 1996; Trombley et al., 2011) and AMPARs (Blakemore and Trombley, 2004) in the OB (See Supplementary Table S1).

Sensory transduction of odor molecules occurs in olfactory sensory neurons (OSNs) and this information is transmitted to cortical structures that encode olfactory discrimination and perception (Trombley and Shepherd, 1993; Ennis et al., 2007, 2015). The first synaptic relay occurs in the OB, whose organization is critical to the processing of odor information. Excitatory and inhibitory amino acid receptors are widely distributed in the OB, with distinct laminar and cellular distributions (Trombley and Shepherd, 1993; Ennis et al., 2007). An understanding of how a neuroactive substance such as zinc modulates the multiple amino acid receptors and ion channels expressed by OB neurons will increase our understanding of the roles that synaptic circuits of OB play in odor information processing and transmission.

Anatomy and Circuitry of The Olfactory Bulb

Odor molecules bind to receptors located on the cilia of OSNs in the olfactory epithelium. Odor binding activates cyclic nucleotide gated channels and the resulting depolarization evokes action potentials that are conducted along the olfactory nerve (ON) to the glomerular layer (GL) of the OB. The OB is divided into multiple layers, consisting of morphologically and functionally distinct types of cells (Ennis et al., 2007, 2015; Nagayama et al., 2014; Kosaka and Kosaka, 2016). See Figure 1 in Corthell et al. (2013) for a diagram illustrating the neuronal populations, layers and synaptic connections in the OB.

The GL is deep to the ON layer (made up of axons of OSNs) and it contains discreet spherical structures known as glomeruli (Golgi, 1875). Within glomeruli, axons of OSNs form glutamatergic axodendritic synapses with mitral and tufted (M/T) cells (Berkowicz et al., 1994; Ennis et al., 1996) as well as intrinsic interneurons (Ennis et al., 2001). M/T cells represent the major output cells in the OB, and these cells are also glutamatergic (Liu et al., 1989; Trombley and Westbrook, 1990; Christie et al., 2001). Three morphologically distinct populations of OB interneurons surround the glomeruli, known collectively as juxtaglomerular (JG) cells: periglomerular (PG) cells, short axon (SA) cells and external tufted (ET) cells (Golgi, 1875; Pinching and Powell, 1971; Shepherd, 1972). JG cells make synaptic contacts with each other as well as with OSN terminals and M/T cells (Shepherd, 1972; Hsia et al., 1999; Berkowicz and Trombley, 2000; Ennis et al., 2001, 2015; Davila et al., 2003; Nagayama et al., 2014; Kosaka and Kosaka, 2016; Liu et al., 2016; Vaaga et al., 2017).

Within the glomeruli, several types of circuits exist (Ennis et al., 2015). M/T cells provide excitatory (glutamatergic) input to PG cells, which provide feedback and feedforward inhibitory input onto M/T cells at reciprocal dendrodendritic GABAergic synapses (Shepherd et al., 2004; Ennis et al., 2007, 2015). This local intraglomerular inhibition is mostly mediated by GABAA receptors (Ennis et al., 2007, 2015). More recent evidence suggests that ET cells play a role in this inhibition. Spontaneous and ON-evoked spike bursts from ET cells drive most PG cells, and this ET cell→PG cell circuit has been shown to result in a major part of the intraglomerular inhibition feeding forward onto M/T cells and feeding back onto ET cells (Ennis et al., 2015). It also has been shown that OSN terminals synapse on ET cells, which, in turn, provide glutamatergic excitatory input to M/T cells and all other JG cells (Hayar et al., 2004a,b, 2005; De Saint Jan et al., 2009; Gire et al., 2012). These findings suggest that ON→M/T cell and ON→ET cell→M/T cell circuits operate in parallel to generate glomerular output (De Saint Jan et al., 2009; Gire et al., 2012; Ennis et al., 2015).

Spanning multiple glomeruli are the dendrites and axons of SA cells, which are thought to regulate interglomerular functions (Aungst et al., 2003; Ennis et al., 2007). Recently, it was shown that GABA and dopamine are co-released from a subset of SA cells, which evokes a temporally biphasic inhibition-excitation response in ET cells (Liu et al., 2013) and reduces OSN glutamate release probability via D2 and GABAB receptor activation (Vaaga et al., 2017). Some data suggest that SA cells also may inhibit mitral cells via actions at GABAA receptors (Liu et al., 2016).

Deep to the GL layer is the external plexiform layer (EPL), which mostly consists of mitral cell and granule cell dendrites that come from the mitral cell layer (MCL) and granule cell layer (GCL), respectively. The EPL also includes the cell bodies and dendrites of tufted cells and heterogeneous subtypes of intrinsic interneurons (Ennis et al., 2015). The EPL contains dendrodendritic synapses, which are reciprocal excitatory/inhibitory synapses between M/T cell and granule cell dendrites (Hirata, 1964; Rall et al., 1966; Price and Powell, 1970b; Woolf et al., 1991). Such synapses provide feedback and lateral inhibition of M/T cells (Ennis et al., 2015). Glutamate released from M/T cells evokes a dual-component response in granule cells, consisting of fast AMPAR-mediated and slower NMDA-receptor-mediated responses (Trombley and Westbrook, 1990; Isaacson and Strowbridge, 1998; Schoppa et al., 1998; Aroniadou-Anderjaska et al., 1999; Chen et al., 2000; Isaacson, 2001). Such activation of granule cells induces GABAA receptor-mediated inhibition of M/T cells (Chen et al., 2000; Isaacson and Vitten, 2003; Dietz and Murthy, 2005). GABA release is coupled to calcium influx via both NMDA receptors (NMDARs) and voltage gated P/Q- and N-type calcium channels (Isaacson and Strowbridge, 1998; Chen et al., 2000; Halabisky et al., 2000; Isaacson, 2001; Egger et al., 2003, 2005).

Below the EPL is the MCL, which contains the somata of mitral cells and granule cells (Ennis et al., 2015). Axon collaterals of mitral cells terminate in the internal plexiform layer (IPL) and GCL (Price and Powell, 1970a; Mori et al., 1983) and also exit the OB as the lateral olfactory tract, which projects to the primary olfactory cortex and other olfactory encoding structures (de Olmos et al., 1978; Ennis et al., 2007, 2015).

Below the MCL is the IPL, which is composed mostly of dendrites of granule cells and axons of M/T cells and centrifugal sources (Ennis et al., 2007). Deep to the IPL is the GCL, which is occupied mostly by granule cells (Nagayama et al., 2014). Granule cells are axon-less inhibitory GABAergic cells whose dendritic spines form reciprocal synapses with secondary dendrites of M/T cells in the EPL (Nagayama et al., 2014). The IPL and GCL also contain a group of diverse interneurons loosely referred to as deep SA cells that include horizontal, Blanes, Golgi and Cajal cells (Eyre et al., 2008).

Zinc Transporters

As well as modulating synaptic transmission, zinc functions as an intracellular signal transducer, which is regulated by zinc transporters (Hara et al., 2017). Zinc transporters can be categorized into two major families: SLC39s/ZIPs and SLC30s/ZnTs (Hojyo and Fukada, 2016; Hara et al., 2017). The ZIP family is responsible for zinc uptake into the cells (Fukada et al., 2014; Hojyo and Fukada, 2016), while the ZnT family reduces the intracellular cytoplasmic zinc content by effluxing zinc from the cytosol or transporting it into vesicles or intracellular organelles (Fukada et al., 2014; Hojyo and Fukada, 2016).

Several studies described here have examined the distribution and/or function of ZnTs. ZnT-1 protects cells against zinc toxicity (Shusterman et al., 2014) and regulates cellular zinc levels (Segal et al., 2004). ZnT-3 is responsible for zinc uptake into synaptic vesicles (Palmiter et al., 1996; Cole et al., 1999), thus, has served as another marker of zinc-enriched terminals. The pattern of ZnT3 mRNA expression in the brain is closely related to the distribution of zinc-containing neurons (McAllister and Dyck, 2017) and most regions of the brains of ZnT3 knockout (KO) mice lack free histochemically reactive zinc (Cole et al., 1999; McAllister and Dyck, 2017).

Locations of Vesicular Zinc

Vesicular zinc is most often contained in subpopulations of glutamatergic neurons (Beaulieu et al., 1992; Slomianka, 1992). Some vesicular zinc is also found in glycinergic and GABAergic neurons in the spinal cord and cerebellum (Birinyi et al., 2001; Wang et al., 2001, 2002). The hippocampus, amygdala, cortex and OB are rich in zinc-containing glutamatergic neurons (Dyck et al., 1993; Frederickson and Moncrieff, 1994; Pérez-Clausell, 1996; Ichinohe and Rockland, 2005). These neurons project internally to sites within the telencephalon (Nakashima and Dyck, 2009), with their pathways mostly found in the cerebral cortex and limbic regions of the forebrain (Bitanihirwe and Cunningham, 2009).

Zinc in the OB is highly concentrated in the GL and GCL (Friedman and Price, 1984; Jo et al., 2000; Sekler et al., 2002; Blakemore et al., 2013). Early histochemical studies demonstrated high concentrations of zinc in the vesicular compartment of the OB (Friedman and Price, 1984; Pérez-Clausell and Danscher, 1985; Masters et al., 1994). In a 2000 study, zinc autometallography (AMG) and ZnT3 immunocytochemistry showed zinc-enriched presynaptic terminals that make contact with granule cells in the GCL and M/T cells and PG cells in the GL (Jo et al., 2000). They suggested that the two sources of zinc-enriched synapses in the mouse OB are centrifugal fibers projecting to granule cells and PG cells and OSN terminals contacting dendrites of M/T cells and PG cells (see Figure 6 in Jo et al., 2000 for a schematic diagram of potential zinc-enriched terminals projecting to the OB). These investigators later used zinc selenium AMG to examine the origins of centrifugal zinc-enriched pathways that project to the rat OB (Mook Jo et al., 2002). They found that the main centrifugal sources of zinc-enriched terminals in the rat OB originate from the nuclei composing the primary olfactory cortex. Collectively, these results suggest that olfactory zinc-enriched terminals are involved in odor information processing.

Another study explored the complementary distributions of chelatable zinc and the zinc transporter ZnT-1 in the mouse OB (Sekler et al., 2002). Zinc, visualized with TSQ histofluorescence, was most heavily concentrated in the olfactory glomeruli, the IPL, and the GCL; negligible zinc was seen in the EPL and MCL. ZnT-1 immunoreactivity was most pronounced in the MCL and in PG cells surrounding the glomeruli, with moderate labeling of granule cells. In brain sections, ZnT-1- immunoreactive neurons in the OB were closely related to synaptic zinc.

As previous studies only characterized zinc’s location in the OB in fixed tissues, we recently conducted a study aimed at advancing these observations by using the zinc-sensitive, fluorescent probe Zinpyr-1 (ZP1; Burdette et al., 2001) to investigate the location and release of zinc in living OB slices (Blakemore et al., 2013). ZP1 is a membrane-permeable probe with a fluorescence turn-on that is selective for ionic zinc, thus, permitting zinc in synaptic vesicles to be viewed (Blakemore et al., 2013).

First, we used ZP1 to visualize the location of vesicular zinc in OB slices from 19 to 21-day-old rats (Blakemore et al., 2013). We found high-intensity ZP1 fluorescence in both the GL (vesicular zinc from OSN terminals and centrifugal fibers) and the GCL (vesicular zinc from centrifugal fibers); only a very weak signal was seen in the EPL, which does not contain much zinc.

To demonstrate that the observed signals reflect free ionic zinc, we applied the zinc chelators tris(2-pyridylmethyl)amine (TPA, 30 μM; Ghosh et al., 2010; Huang et al., 2013) and N,N,N′,N′-tetrakis (2-pyridylmethyl)ethylenediamine (TPEN, 30 μM) to absorb zinc ions from ZP1-treated slices. Incubation with either chelator produced similar reductions in the ZP1 signal, suggesting that the fluorescence emission from ZP1 denotes binding to ionic zinc (Blakemore et al., 2013). Incubation of the ZP1-treated OB slices in a high (50 mM) extracellular potassium solution also rapidly eliminated the zinc signal, likely by causing the release of zinc-containing synaptic vesicles due to depolarization. Together, these results obtained with the ZP1 probe and zinc chelators demonstrated, for the first time in living OB slices, that vesicular zinc is mostly localized within the GL and GCL.

Release of Vesicular Zinc

The co-localization of zinc and glutamate in a subset of zinc-enriched neurons, along with evidence that zinc is released from synaptic vesicles in response to depolarizing stimuli, suggests that zinc is released during synaptic transmission. In early (1980s) experiments, the release of zinc into the synaptic cleft was first demonstrated in vitro in hippocampal slice preparations following depolarization in response to electrical (Howell et al., 1984) or chemical (high potassium or kainate; Assaf and Chung, 1984) stimuli. This zinc release was both calcium- and depolarization-dependent (Assaf and Chung, 1984; Howell et al., 1984), consistent with a vesicular release mechanism.

Various tools for studying vesicular zinc have recently emerged. These include highly selective fluorescent probes that can be used to detect mobile zinc in live cells and tissues, different types of KO mice (e.g., ZnT3 KO mice lacking vesicular zinc; Palmiter et al., 1996; Cole et al., 1999), and various zinc chelators (Goldberg et al., 2016).

Zinc imaging combined with electrophysiological recordings has provided additional direct evidence of vesicular zinc release (Vogt et al., 2000; Li et al., 2001; Molnár and Nadler, 2001; Ueno et al., 2002; Qian and Noebels, 2005, 2006; Paoletti et al., 2009). Using microfluorescence imaging in rat hippocampal slices, Li et al. (2001) showed that electrical stimulation of mossy fibers (MFs) caused the immediate release of zinc from synaptic terminals into the extracellular microenvironment. The membrane-impermeable, zinc-selective fluorescent dye FluoZin-3 (Gee et al., 2002) was used to demonstrate zinc release during exocytosis evoked by action potentials in hippocampal slices at zinc-enriched MF (Qian and Noebels, 2005) and CA3–CA1 (Qian and Noebels, 2006) synapses. The absence of zinc signal observed in slices from ZnT3 KO mice indicated that the released zinc came from synaptic vesicles (Qian and Noebels, 2005, 2006).

Another study in the rat hippocampus used the selective membrane-impermeable zinc chelator ZX1 (Burdette et al., 2001; Chang and Lippard, 2006; Zhang et al., 2007) to show that vesicular zinc released in response to electrophysiological stimulation modulates long-term potentiation (LTP) at MF-CA3 synapses (Pan et al., 2011). A subsequent study used genetic tools (e.g., KO and knockin mice), the fast chelating agent tricine (Paoletti et al., 1997), and electrophysiological recordings to demonstrate that vesicular zinc released via exocytosis modulates activity at postsynaptic NMDARs at the MF-CA3 and Schaffer collateral-CA1 synapses (Vergnano et al., 2014).

In a series of experiments performed in the zinc-enriched dorsal cochlear nucleus (DCN), the effects of both synaptic and tonic (nonsynaptic) zinc were explored. In one study, ZX1, electrophysiological methods, and glutamate uncaging were used to demonstrate that action potential-evoked release of synaptic zinc was required to inhibit alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor-mediated currents (Kalappa et al., 2015). In another study, chelation of zinc with ZX1 increased the frequency of action potentials in the DCN. This is consistent with their interpretation that tonic zinc inhibits spontaneous action potentials in the DCN by a combination of glycine receptor potentiation and an increase in glycine release probability (Perez-Rosello et al., 2015). Anderson et al. (2015) also used ZX1 and a novel extracelluar ratiometric fluorescent zinc sensor (LZ9) to show that both synaptic and tonic zinc modulate extrasynaptic NMDARs in the DCN. Finally, Perez-Rosello et al. (2013) found that synaptic zinc in the DCN reduces the probability of transmitter (glutamate) release by acting with the putative metabotropic zinc-sensing receptor GPR39 to promote endocannabinoid synthesis.

Recently, we used zinc imaging combined with electrophysiology to explore the release of synaptic zinc in the OB (Blakemore et al., 2013). We incubated OB slices from 19 to 21-day-old rats with ZP1 and induced zinc release by electrically stimulating the ON layer. Placement of the stimulating electrode was at the ON-GL boundary to permit isolation of a single or several glomeruli innervated by the adjacent bundle of ON axons. The ON axon bundle was electrically stimulated with a pattern designed to mimic the breathing cycle in rats (Youngentob et al., 1987). These stimulation patterns selectively decreased the zinc signal in glomeruli adjacent to the stimulating electrode, whereas the zinc signal in nearby glomeruli remained unchanged. Thus, we have presented the first evidence that OSN terminals within individual glomeruli release vesicular zinc in response to patterned electrical stimulation of the ON simulating the rat’s breathing cycle. This observation is important given that the GL expresses a high density of neurotransmitter receptor and ion channel targets for modulation by zinc co-released with glutamate from OSNs.

The exact quantity of vesicular zinc released into the synaptic cleft is unclear. The concentration of zinc in the synaptic cleft after an excitatory event is influenced by factors such as the number of zinc vesicles that fuse to the presynaptic membrane (Goldberg et al., 2016). Early studies suggested that extracellular zinc concentrations ranged from 10 to 300 μM following depolarization of neurons (Frederickson et al., 1983; Assaf and Chung, 1984; Howell et al., 1984; Xie and Smart, 1991). However, synaptic concentrations of zinc could be greater given that the synaptic cleft comprises only part of the extracellular space. More recently, it has been estimated that peak zinc concentrations in the synaptic cleft following exocytosis range from 10 nM to >100 μM (Vogt et al., 2000; Ueno et al., 2002; Paoletti et al., 2009; Vergnano et al., 2014; Goldberg et al., 2016). Synaptic concentrations of zinc in the micromolar range are not unexpected since the synaptic concentrations of glutamate (Clements et al., 1992), GABA (Jones and Westbrook, 1995), and glycine (Beato, 2008) all exceed 1 mM. This supports the notion that vesicular release of zinc generates synaptic concentrations well within a range to effectively alter ion channel function.

Zinc Modulation of Amino Acid Receptors and Synaptic Transmission with Focus on the OB

A number of studies suggest that zinc modulates amino acid receptors and synaptic transmission (Smart et al., 1994; Bitanihirwe and Cunningham, 2009; Nakashima and Dyck, 2009; Paoletti et al., 2009). This section examines the effects of zinc on several types of excitatory amino acid receptors (NMDARs, AMPARs) and inhibitory amino acid receptors (GABAA receptors, glycine receptors), with special emphasis on zinc’s actions in the OB.

Zinc’s Effects on NMDA Receptors

Much attention has been focused on the effects of zinc on NMDARs given their role in neuronal excitability and plasticity (Nakashima and Dyck, 2009). Early studies showed potent inhibition of NMDARs by zinc at a range of zinc concentrations (1 μM–1 mM; Smart et al., 1994). At low micromolar concentrations, zinc caused voltage-independent non-competitive inhibition of NMDAR-mediated responses in hippocampal and cortical neurons by decreasing open channel probability (Peters et al., 1987; Westbrook and Mayer, 1987; Mayer et al., 1989; Legendre and Westbrook, 1990; Xie et al., 1993; Paoletti et al., 2009). At higher concentrations, zinc bound to the magnesium site within the channel causing voltage-dependent inhibition of NMDAR-mediated currents (Paoletti et al., 2009).

NMDARs are formed by multiple subunits and zinc’s effects on NMDARs vary with subunit composition. In 2009, new nomenclature to describe these subunits was introduced. The seven different subunits described consist of the GluN1 subunit, four different GluN2 subunits (GluN2A, GluN2B, GluN2C and GluN2D), and two GluN3 subunits (GluN3A and GluN3B; see Table 3 in Collingridge et al., 2009). Here, to avoid confusion, we use the nomenclature used in previously published articles with the new nomenclature indicated parenthetically.

Whereas zinc is a specific antagonist of some GluN2-containing NMDARs (Paoletti et al., 2013), differences in zinc affinity for various GluN2 subunits exist. Zinc causes a high affinity voltage-independent inhibition of NR1/NR2A (GluN1/GluN2A)-containing receptors, and a low affinity inhibition of NR1/NR2B (GluN1/GluN2B)-containing receptors (Chen et al., 1997; Paoletti et al., 1997; Low et al., 2000; Rachline et al., 2005). The functional consequence of this difference in zinc affinity is that low (nanomolar) zinc concentrations only inhibit GluN2A-containing receptors, whereas higher (micromolar) zinc concentrations are needed to inhibit GluN2B-containing receptors (Paoletti et al., 1997, 2013; Rachline et al., 2005).

Early autoradiographic, immunocytochemical, in situ hybridization and electrophysiological studies demonstrated the presence of NMDARs in the OB (Trombley and Shepherd, 1993). NMDARs are found throughout the OB aside from the IPL (Watanabe et al., 1993; Monyer et al., 1994; Petralia et al., 1994; Ennis et al., 2007).

We used whole-cell recording techniques to examine the effects of zinc on NMDARs expressed by OB neurons. We found that zinc was an effective antagonist of NMDA-mediated currents in rat OB neurons in primary culture and acutely isolated from adult animals (Trombley and Shepherd, 1996). Whereas 1 μM zinc had little effect on currents evoked by 100 μM NMDA, 100 μM zinc almost completed blocked these currents. The half maximal inhibitory concentration (IC50) was 19 μM (Trombley and Shepherd, 1996).

In our 1998 study, we investigated zinc’s effect on spontaneous, glutamate-mediated excitatory activity in the rat OB, mediated in part by NMDARs (as well as AMPA and kainate receptors; Trombley et al., 1998). At 100 μM, zinc caused a transition from asynchronous spiking to synchronous busting while also reducing synaptic frequency. Our subsequent study showed similar inhibitory effects of zinc (100 μM) on spontaneous excitatory activity (Horning and Trombley, 2001).

Because glutamate and zinc are co-localized in OSN terminals, there could be a variety of outcomes in regard to modulation of odor processing depending on their patterns of synaptic release. Both NMDA and AMPARs are involved in the activation of OB neurons by OSNs (Berkowicz et al., 1994; Ennis et al., 1996, 2001; Aroniadou-Anderjaska et al., 1997; Chen and Shepherd, 1997). Thus, inhibition of NMDARs by zinc co-released with glutamate from OSN terminals may modify excitatory transmission at these synapses. The NMDAR-mediated component of the response of M/T cells to glutamate released from OSNs is unusually long, resulting in a late spiking component (Ennis et al., 1996, 2001; Aroniadou-Anderjaska et al., 1997). Zinc may attenuate this NMDAR-mediated component of the response to ON input, thus, alter the spiking rate.

Zinc’s Effects on AMPARs

The AMPA-subtype of glutamate receptors mediates most fast excitatory transmission in the CNS. AMPARs play an important role in normal function and plasticity of the brain. Zinc often had biphasic, concentration-dependent effects at AMPARs expressed in Xenopus oocytes (Rassendren et al., 1990) or on neurons from various brain regions (Mayer et al., 1989; Bresink et al., 1996). Low concentrations of zinc (50–300 μM) have been shown to potentiate AMPAR-mediated currents, while higher concentrations (1–3 mM) inhibit these currents (Mayer et al., 1989; Bresink et al., 1996). Furthermore, zinc-sensitive and zinc-insensitive AMPARs have been shown to co-exist in some brain regions (Mayer et al., 1989; Lin et al., 2001).

Early autoradiographic, immunocytochemical, in situ hybridization, and electrophysiological studies demonstrated the presence of AMPARs in the OB (Trombley and Shepherd, 1993). Apart from the ONL and subependymal layer, AMPARs are distributed throughout the OB (Petralia and Wenthold, 1992; Martin et al., 1993; Molnár et al., 1993; van den Pol, 1995; Ennis et al., 2007). AMPARs are found on the cell bodies and dendrites of mitral, tufted and JG cells (Petralia and Wenthold, 1992; Molnár et al., 1993; Giustetto et al., 1997; Montague and Greer, 1999). Because zinc and glutamate are co-localized in synaptic vesicles within the OB, we used whole-cell recording techniques to examine the effects of zinc on AMPARs on cultured rat OB neurons (Blakemore and Trombley, 2004). We found that the effects of various concentrations of zinc on AMPA (50 μM)-evoked currents mediated by OB AMPARs can be biphasic, uniphasic, or absent.

In contrast to previous studies in other brain regions, we analyzed the effects of multiple concentrations of zinc (30 μM, 100 μM and 1 mM) on AMPARs in the same cell in a subpopulation of OB neurons (N = 47; Blakemore and Trombley, 2004). We observed the expected biphasic response to zinc, current potentiation at 30 μM and/or 100 μM and inhibition at 1 mM, in ~46% of M/T cells compared with only ~16% interneurons. This difference in the frequency of a biphasic response between cell types was statistically significant. AMPARs on no M/T cells and two interneurons were insensitive to all three zinc concentrations. Uniphasic potentiating-only or inhibiting-only responses were observed at AMPARs on the remaining cells. Thus, the range of zinc concentrations we used was more likely to generate uniphasic rather than biphasic effects at AMPARs in the OB.

AMPARs are formed from four subunits: GluR1–4 (Dingledine et al., 1999). In 2009, new nomenclature to describe these subunits was introduced (see Table 3 in Collingridge et al., 2009), which re-named GluR1–4 as GluA1–4. Here, to avoid confusion, we use the nomenclature used in previously published articles with the new nomenclature indicated parenthetically. Zinc’s effects on AMPARs vary with the receptor’s subunit composition (Dreixler and Leonard, 1994), extracellular environment (Dreixler and Leonard, 1997) and alternative splicing of the AMPAR gene (Shen and Yang, 1999). In Xenopus oocytes, low concentrations of zinc (4–7.5 μM) augmented currents mediated by homomeric GluR3 (GluA3), but not homomeric GluR1 (GluA1) receptors, suggesting the GluR3 (GluA3) subunit is necessary for zinc potentiation (Dreixler and Leonard, 1994). Heteromeric expression of GluR3 (GluA3) with GluR2 (GluA2) and GluR1 (GluA1) with GluR2 (GluA2) both resulted in the absence of zinc potentiation (Dreixler and Leonard, 1994). Dreixler and Leonard later suggested that the zinc potentiation site in AMPARs may encompass more than one GluR3 (GluA3) subunit and that this site may be altered when the GluR2 (GluA2) subunit is included in the receptor (Dreixler and Leonard, 1997). The responsiveness of AMPARs to zinc is also affected by alternative splicing of the AMPAR gene, with resistance to modulation by zinc seen in flop variants but not flip variants (Shen and Yang, 1999).

Given these findings, one explanation for the differences in zinc sensitivity among OB neurons observed in our 2004 study is that AMPAR subunits are heterogeneously distributed within the OB. Support for this stems from immunohistochemical data showing that all AMPAR subunit proteins are expressed in the OB but follow distinct laminar, cellular and subcellular distributions (Montague and Greer, 1999). Our previous finding that the kinetics of OB AMPARs vary within and between OB neuron subtypes (Blakemore and Trombley, 2003) further suggests the diverse expression of AMPAR subunits and products of alternative splicing of the AMPAR gene (Dingledine et al., 1999).

In our 2003 study of AMPAR kinetics, we observed differences in the rate and extent of receptor desensitization between M/T cells and interneurons in primary culture (Blakemore and Trombley, 2003). We also observed differences between M/T cells and interneurons in their AMPARs’ sensitivity to cyclothiazide (Blakemore and Trombley, 2003), a drug that reduces desensitization of flip splice variants more extensively than flop splice variants (Dingledine et al., 1999). Our results suggest that AMPARs on interneurons contain a relatively higher percentage of flop subunits, whereas those on M/T cells contain a relatively higher percentage of flip subunits (Blakemore and Trombley, 2003). Given the results of Shen and Yang (1999), this may explain why AMPARs on OB interneurons were less likely than AMPARs on M/T cells to show a biphasic response to zinc (Blakemore and Trombley, 2004). It also has been reported that application of cyclothiazide abolishes zinc-mediated potentiation of currents evoked by glutamate, suggesting that zinc may potentiate AMPAR-receptor mediated currents by inhibiting receptor desensitization (Lin et al., 2001).

Evidence from our molecular biology study (Horning et al., 2004) provides support for the conclusions from our kinetic study that the OB has a diverse expression of AMPAR subunits and splice variants. In this study, we found that while the rat OB expresses mRNAs for GluR1–4 (GluA1–4), the relative amounts differ i.e., GluR2 > GluR1 >> GluR4 >> GluR3. Similar amounts of both variants were found in GluR2 and GluR4 transcripts (GluR2: ~55% flip, ~45% flop; GluR4: ~60% flip, ~40% flop). However, GluR1 and GluR3 transcripts consisted mostly of flip (GluR1: ~95% flip, ~5% flop; GluR3: ~93% flip, ~7% flop; Horning et al., 2004). The relatively higher levels of the flop variant for the GluA2 subunit may suggest that interneurons are more likely to contain GluA2, thus, are less likely to be sensitive to zinc.

Our 2004 study also showed that the frequency and magnitude of zinc’s effects at AMPARs on OB neurons vary with cell type. Modulation of AMPAR-mediated currents was observed in a greater percentage of M/T cells than interneurons at low (30 μM, 100 μM) but not high (1 mM) zinc concentrations. However, zinc’s effect on AMPAR-mediated currents was greater in magnitude in interneurons than in M/T cells (Blakemore and Trombley, 2004). AMPARs that lack (or contain unedited) GluR2 (GluA2) are also highly permeable to calcium (Hume et al., 1991; Verdoorn et al., 1991; Burnashev et al., 1995) and evidence suggests that OB interneurons primarily express AMPARs with low Ca2+ permeability (Jardemark et al., 1997). Thus, our observation that zinc modulated AMPARs on interneurons less often than AMPARs on M/T cells again may reflect the fact that AMPARs on interneurons have a higher GluA2 content.

In later studies, we (Blakemore et al., 2006) and others (Ma and Lowe, 2007; Pimentel and Margrie, 2008) demonstrated that a subset of AMPARs activated by OSNs, as well as AMPARs activated between M/T cell dendrites, are calcium permeable. This finding may be important, as calcium-permeable AMPA/kainate receptors on cortical and hippocampal neurons have been shown to also flux zinc, with a myriad of effects in the postsynaptic neuron (Yin and Weiss, 1995; Weiss and Sensi, 2000). Once within the cell, zinc can trigger widespread disruptions of normal cellular functions, including disruption of calcium homeostasis and tubulin assembly, inhibition of mitochondrial electron transport, over-activation of calcium-mediated enzymes, and the production of reactive oxygen species (Kress et al., 1981; Csermely et al., 1988; Choi and Koh, 1998; Weiss and Sensi, 2000). Zinc influx through calcium-permeable AMPA/kainate receptors also may be involved in intracellular zinc signaling (Takeda et al., 2007, 2009; Tamano and Takeda, 2011).

Calcium entry into cells via calcium-permeable AMPARs also may influence transmitter release. In the retina, calcium-permeable AMPARs, which are activated by glutamate released from rod bipolar cells, provide the influx of calcium needed to cause GABA release from A17 amacrine cells (Chávez et al., 2006). In the OB, calcium-permeable AMPARs have been shown to be the major receptor type underlying glutamatergic signaling in neural precursor cells (NPCs; Darcy and Isaacson, 2010). Under conditions of decreased transmitter uptake, glutamate spillover from synapses in the GCL can activate nonsynaptic NPC AMPARs and increase intracellular calcium levels (Darcy and Isaacson, 2010).

Recently, we showed that zinc potentiates AMPAR-mediated synaptic events (Blakemore et al., 2013). Using whole-cell current-clamp and voltage-clamp recording, we examined the effects of zinc on isolated AMPAR-mediated excitatory post-synaptic potentials (EPSPs) and post-synaptic currents (EPSCs) recorded from M/T cells and interneurons in primary cultures of rat OB neurons. Co-application of 100 μM zinc potentiated AMPAR-mediated EPSPs and/or EPSCs in 60% of M/T cells compared with only 23% of interneurons (Blakemore et al., 2013). Thus, consistent with our 2004 study, the frequency of zinc-mediated potentiation differed between cell types, which is likely due to variability in the AMPAR subunit composition. Also, consistent with our 2004 study, zinc perfusion reduced both EPSPs and EPSCs in subsets of M/T cells (10%) and interneurons (30%) and appeared to have no measureable effect in some M/T cells (30%) and interneurons (46%; Blakemore et al., 2013), further supporting the co-existence of cells with zinc-sensitive and zinc-insensitive AMPARs in the OB.

Together, these results have important implications for synaptic transmission, hence, odor processing. For example, zinc-mediated potentiation of the relatively small and brief AMPAR-mediated current could increase the typically minor role AMPARs play in reciprocal inhibition between M/T cells and interneurons (Isaacson and Strowbridge, 1998; Schoppa et al., 1998). In addition, we propose that calcium influx into cells through calcium-permeable AMPARs may be important at hyperpolarized potentials when the driving force for calcium entry is large but calcium flux via NMDARs is blocked by magnesium. Zinc-mediated potentiation of currents mediated by calcium-permeable AMPARs would increase calcium flux, potentially enhancing transmitter release.

Mitral and tufted cells within a glomerulus are synchronized by patterns of glutamate release from OSN (Carlson et al., 2000; Schoppa and Westbrook, 2001, 2002; Christie and Westbrook, 2006; Schoppa, 2006). This synchronization depends partly on AMPAR activation (Schoppa and Westbrook, 2002), generating oscillations with temporal and spatial features in response to odors. Glutamate release evokes the spatial patterns, but GABAergic neurons regulate the temporal patterns (Schoppa, 2006) that together underlie OB encoding of olfactory information. The results of our 2013 study extend our previous 2004 analyses of synaptic/extrasynaptic AMPARs (Blakemore and Trombley, 2004) by isolating the effects of zinc on synaptically activated AMPARs important to this OB circuit function. In M/T cells, AMPAR-mediated EPSPs were potentiated by zinc, thus, enhancing the frequency of synaptic summation (Blakemore et al., 2013). Such enhancement would facilitate neuronal synchronization and modify information processing by the OB.

Zinc’s Effects on GABAA Receptors

GABAA receptors are ligand-gated ion channel receptors that flux chloride ions and mediate fast inhibitory transmission (Schofield et al., 1987; Barnard et al., 1988). Early electrophysiological evidence suggested the presence of functional GABAA receptors in the OB (Trombley and Shepherd, 1993), and a variety of studies have shown GABAA receptors to be expressed throughout the OB aside from the IPL (Laurie et al., 1992; Panzanelli et al., 2005; Ennis et al., 2007).

Westbrook and Mayer (1987) showed that zinc (5 μM) inhibited the response to GABA in cultured hippocampal neurons in a voltage-insensitive manner (Westbrook and Mayer, 1987). A variety of subsequent studies found that exogenously applied zinc mostly (but not always) inhibited GABAA receptors (Smart et al., 1994).

Wang et al. (2001, 2002) demonstrated the presence of zinc-enriched GABAergic terminals in subpopulations of neurons in mouse spinal cord and cerebellar cortex. Ruiz et al. (2004) showed that both GABA and zinc are contained in the same hippocampal MF varicosities. Their finding that exogenously applied zinc depressed IPSCs recorded in hippocampal slices led them to conclude that endogenously released zinc modulates GABAA-receptor-mediated responses at zinc-containing MF synapses. Together, these results suggest that endogenous zinc may modulate inhibitory transmission.

We and others have demonstrated zinc-mediated modulation of GABA-mediated currents in OB neurons. Serafini et al. (1995) reported that GABAA-activated currents recorded in embryonic rat OB cells were partially inhibited by a wide range of zinc concentrations (10–1000 μM). We subsequently reported that zinc (100 μM) was an effective antagonist of GABAA-mediated currents in cultured rat OB neurons; the IC50 was 17 μM (Trombley and Shepherd, 1996). Consistent with zinc’s effect on exogenously evoked GABAA-mediated currents, we later showed that 100 μM zinc completely blocked GABA-mediated spontaneous inhibitory transmission (IPSPs) in OB neurons (Horning and Trombley, 2001).

Together, these results suggest that zinc modulates currents mediated by GABAA receptors expressed by OB neurons. As one source of zinc-enriched terminals in the OB is OSN terminals contacting dendrites of M/T cells and PG cells (Jo et al., 2000; Blakemore et al., 2013), zinc released at these synapses may influence inhibition mediated by GABAergic JG cells.

Centrifugal pathway inputs from a variety of brain regions affect odor information processing in the OB. A second source of zinc-containing synaptic terminals in the mouse OB is centrifugal fibers projecting to granule cells and PG cells. Thus, zinc released from these centrifugal fibers may influence excitatory transmission at these synapses and subsequent inhibitory transmission mediated by GABAergic granule cells and PG cells.

Zinc’s Effects on Glycine Receptors

The glycine receptor is a membrane-spanning protein that contains a Cl-selective ion channel, producing an inhibitory response when glycine binds that is similar to GABAergic inhibition. In the OB, immunoreactivity for glycine has been demonstrated in EPL and in granule cells (Pourcho et al., 1992), and robust labeling for glycine receptors exists (van den Pol and Gorcs, 1988; Weltzien et al., 2012). High levels of glycine are also found in the OB (Ennis et al., 2007), which inhibit rat OB cells (Trombley and Shepherd, 1994; Trombley et al., 1999).

Zinc is reportedly co-localized with glycine in inhibitory neurons in the lamprey and mouse spinal cord (Birinyi et al., 2001; Wang et al., 2001), and additional studies suggest that synaptic zinc is essential for the proper in vivo functioning of glycinergic neurotransmission in the spinal cord and brainstem (Hirzel et al., 2006). In rat spinal cord, low zinc concentrations (20 nM–1 μM) enhanced currents evoked by 100 μM glycine, while higher zinc concentrations (20–50 μM) caused a reversal of this potentiation, followed by progressive current inhibition (Bloomenthal et al., 1994).

Subsequent studies, including our own, have shown similar biphasic (potentiating and inhibiting), concentration-dependent effects of zinc on both native and recombinant glycine receptors (Laube et al., 1995; Trombley and Shepherd, 1996; Lynch et al., 1998; Harvey et al., 1999; Lynch, 2004; Trombley et al., 2011). Zinc’s potentiating and inhibiting concentrations are often reported as <10 μM and >10 μM, respectively (Lynch, 2004). However, the glycine receptor’s subunit composition introduces variability in zinc’s concentration-dependent effects (Miller et al., 2005a), and we have observed biphasic effects in the OB at higher concentrations of zinc (Trombley and Shepherd, 1996; Trombley et al., 2011).

Results from our 1996 study advance the results of previous studies by suggesting that the effects of zinc on glycine receptors in the rat OB only occur with use of low concentrations of glycine, which do not desensitize the receptors. We showed that 100 μM zinc potentiated the current when a non-desensitizing concentration of glycine (30 μM) was used. A higher concentration of zinc (1 mM) inhibited currents evoked by 30 μM glycine. However, zinc (up to 1 mM) had no effect on the steady-state, desensitized component of the current evoked by a high concentration of glycine (300 μM; Trombley and Shepherd, 1996).

Because the synaptic cleft concentration of glycine is greater than 2 mM (Beato, 2008), glycine receptors are rapidly saturated and transition to a desensitized state during synaptic transmission. With low mM concentrations, glycine receptors progressively desensitize with even short (1 ms) pulses of glycine when applied in excess of 1 Hz (Rigo and Legendre, 2006). Of additional significance, it has been recently demonstrated that pulse applications, within a normal physiological frequency, cause significant receptor desensitization and that much of this happens in between the pulses when glycine is absent (Papke et al., 2011). Consequently, glycine receptors undergo significant desensitization during normal physiological conditions.

Zinc’s inhibition of glycine receptors is reportedly due to stabilization of the closed state of the receptor (Lynch, 2004). Therefore, we hypothesized that high micromolar glycine concentrations may result in the occupation of all glycine binding sites and a conformational change in the receptor (to a desensitized state) that prevents zinc binding or the effects of bound zinc. To test this hypothesis, we continuously applied 300 μM zinc during intermittent pulse application of 300 μM glycine; we found that such pre-application of zinc slowly inhibited currents evoked by desensitizing concentrations of glycine recorded in cultured rat OB neurons (Trombley et al., 2011). Another protocol revealed increasing zinc-mediated inhibition of this current with increasing durations of zinc exposure (Trombley et al., 2011).

Several types of receptors and ion channels allow synaptically released zinc to enter the postsynaptic neuron (Koh and Choi, 1994; Sensi et al., 1997, 1999; Kerchner et al., 2000; Weiss and Sensi, 2000; Tamano and Takeda, 2011), where it could interact with an intracellular domain of the glycine receptor. To explore this hypothesis, we compared the fluorescence of neurons loaded with a zinc-sensitive probe (Newport Green) under control conditions (no added zinc) to conditions in which 300 μM extracellular zinc was applied for 60 s. Compared with control neurons, neurons exposed to 300 μM zinc were more robustly labeled, confirming that extracellular application of zinc results in elevated intracellular zinc (Trombley et al., 2011).

To further explore our hypothesis that intracellular zinc contributes to inhibition, we included several types of zinc chelators (10 mM EDTA, 100 μM dipicolinic acid, or 100 μM 1,10-phenanthroline) in our recording electrode and evoked a glycine current with application of 300 μM glycine; 300 μM zinc was then preapplied for 60 s and another glycine-evoked current was measured. None of the chelators prevented the inhibitory effects of externally applied zinc on the current. These results may suggest that the chelation process was too slow to prevent an intracellular effect of zinc or that the increase in intracellular zinc did not affect inhibition. However, the most likely explanation for this finding is that an extracellular binding site mediates inhibition by externally applied zinc (Trombley et al., 2011). Consistent with this notion, the site(s) for zinc’s inhibitory actions on the glycine receptor appear(s) appear to be located on the external face of the alpha subunit (Laube et al., 1995, 2000; Nevin et al., 2003; Miller et al., 2005a).

To further examine zinc’s possible intracellular effects, a range of zinc concentrations (1, 10, 100 μM) was added to the recording electrode (Trombley et al., 2011). All concentrations of zinc rapidly increased the amplitude of the current evoked by 300 μM glycine, which was followed by a slight decrease in the current. Inclusion of a chelator with zinc in the electrode blocked these effects of intracellular zinc. These results represent the first report that glycine receptor-mediated currents can be potentiated by low concentrations of intracellular zinc.

Previous studies have suggested that the zinc potentiation site(s) on the glycine receptor is/are located on the external face of the alpha subunit’s N-terminal domain (Laube et al., 1995, 2000; Lynch, 2004; Miller et al., 2005b). However, our finding that low concentrations of intracellular zinc potentiate glycine receptor-mediated currents also suggest the presence on an internal zinc binding site that mediates potentiation by zinc. Several studies have shown that the glycine receptor contains several amino acids (His, Glu, Asp) on the intracellular loops capable of binding zinc (Laube et al., 2002; Lynch, 2004; Miller et al., 2005b).

Together, these results illustrate the complexity of zinc’s effects on glycine receptors as well as the many interacting factors that influence glycinergic inhibition; these factors include the receptor subunit composition, the state of the receptor, the timing of exposure to zinc and glycine, the concentration of zinc exposure, and whether zinc gains intracellular access. Each of these factors would significantly alter the efficacy of glycinergic transmission in the OB.

Zinc Modulation of Voltage-Gated Ion Channels with A Focus on the OB

Zinc modulation of voltage-gated sodium, potassium and calcium channels has been shown by us and others. The below discussion focuses on several types of voltage-gated ion channels found in the OB.

Zinc’s Effects on Voltage-Gated Calcium Channels

Voltage-gated calcium channels are ion channels found in the membranes of excitable cells that are primarily permeable to calcium. Several different types of high-voltage-activated (HVA) calcium channels (L-type, N-type, P/Q-type) exist, which are variously distributed in the central and peripheral nervous systems. The T-type calcium channel is a low-voltage-activated (LVA) channel, which is also expressed by neurons in some brain regions (Mathie et al., 2006).

At physiologic or resting membrane potentials, voltage-gated calcium channels are usually closed. However, at depolarized membrane potentials they open, permitting calcium to enter the cell and triggering the release of neurotransmitter into the synaptic cleft. Voltage-gated and ligand-gated (NMDA and AMPA) calcium channels are also involved in intracellular calcium signaling (Weiss and Sensi, 2000; Foster, 2007), which is influenced by zinc (Tamano and Takeda, 2011).

Early studies demonstrated inhibitory effects of zinc on voltage-gated calcium currents. Zinc was shown to block voltage-gated calcium-channel currents in cultured rat dorsal root ganglion cells (Büsselberg et al., 1994). A later study examined the effects of zinc on voltage-gated calcium channels in acutely dissociated neurons from the horizontal limb of the diagonal band of Broca (hDBB; Easaw et al., 1999), a region enriched in zinc-positive fibers (Mandava et al., 1993). Zinc (50 μM) reversibly attenuated calcium-channel currents recorded in rat hDBB neurons by approximately 85%, with varying effects on L-, N- and P-type HVA calcium-channel currents (Easaw et al., 1999). Zinc (6–200 μM) was later shown to block total HVA calcium-channel currents in pyramidal neurons acutely dissociated from rat piriform cortex, with inhibition of all pharmacological components (L-, N-, P/Q- and R-type) of these currents (Magistretti et al., 2003).

In regard to zinc modulation, Mathie et al. (2006) claimed that the T-type or LVA calcium channels are arguably the most important group. Three channel isoforms (Cav3.1, Cav3.2 and Cav3.3) comprise this family, and neurons in some regions of the OB express these channels (Talley et al., 1999; Liu and Shipley, 2008b; Johnston and Delaney, 2010). Takahashi and Akaike (1991) showed zinc-mediated inhibition (IC50 of 20 μM) of the T-type calcium channel in CA1 pyramidal cells from rat hippocampus. A later study showed zinc-mediated inhibition of T-type calcium channels in rat dorsal root ganglion cells, with 20 μM zinc generating a block exceeding 80% (Büsselberg et al., 1992). The IC50 for zinc inhibition of N- and L-type channels was higher (69 μM). Another study found that zinc blocked 3 types of recombinant T-type calcium channels in a concentration-dependent manner, with a higher affinity for Cav3.2 channels (IC50 2.3 μM; Jeong et al., 2003).

Evidence from a variety of studies suggests that zinc modulation of calcium channels affects calcium influx and transmitter release. A study in rat dorsal root ganglion neurons found that low concentrations (10–100 nM) of zinc evoked an extracellular calcium influx through L-, N- and T-type voltage-gated calcium channels, thus, increasing the release of substance P; higher concentrations (1–100 μM) of zinc attenuated or completely masked these responses (Tang et al., 2009). In the hippocampus and amygdala, high-frequency stimulation with the membrane-impermeable zinc chelator CaEDTA present enhanced exocytosis (Minami et al., 2006; Takeda et al., 2007, 2010). Also in the hippocampus, application of the zinc chelator TPEN during LTP enhanced presynaptic calcium signals without affecting synaptic transmission (Quinta-Ferreira and Matias, 2004). Thus, it has been proposed that endogenous released zinc may control calcium influx through voltage-gated calcium channels (Tamano and Takeda, 2011). Furthermore, voltage-gated calcium channels and calcium-permeable AMPA/kainate receptors allow released zinc to be taken up into the presynaptic and postsynaptic neurons, where it may further modulate calcium influx and serve intracellular zinc signaling functions (Tamano and Takeda, 2011).

In our 2001 study (Horning and Trombley, 2001), we examined the effects of zinc on voltage-gated calcium-channel currents recorded from cultured rat OB neurons. We hypothesized that the effects of zinc on calcium-channel currents could contribute to zinc’s inhibitory effects on excitatory transmission (Trombley et al., 1998; Horning and Trombley, 2001). We induced calcium-channel currents with kinetic profiles typical of HVA channels in M/T cells. Zinc (100 μM) significantly inhibited these calcium-channel currents by about 60%.

Various investigators have reported the presence and actions of T-type channels in the OB. Pignatelli et al. (2005) demonstrated that an LVA (T-type) calcium channel current plays a role in the autorhythmicity of mouse OB dopaminergic neurons in the GL. The effects of zinc released from OSN terminals and centrifugal fibers in the GL on this current could alter the role of these dopaminergic neurons in glomerular circuits.

Mitral cells express LVA Cav3.3 (T-type) channels on their distal apical tuft dendrites (Johnston and Delaney, 2010). In the mouse OB, calcium influx from these channels influences both asynchronous glutamate release from these dendrites and feedback inhibition from PG cells (Fekete et al., 2014). Thus, zinc released from OSN terminals and centrifugal fibers could inhibit these calcium channels and alter reciprocal inhibition at mitral cell-PG cell synapses.

ET cells have a prominent calcium conductance that is activated at approximately −50 mV and has biophysical and pharmacological features of both T- and L-type calcium currents (Liu and Shipley, 2008b). These investigators also demonstrated that L- and/or T-type-mediated LVA calcium currents amplified both subthreshold and suprathreshold synaptic responses in ET cells in response to ON input in mouse OB (Liu and Shipley, 2008a). Again, zinc released from OSN terminals and centrifugal fibers could modify these responses.

Our 2001 results, suggesting that micromolar concentrations of zinc inhibit HVA calcium channels, also may have important implications to bulb function. In the OB, voltage-gated calcium (P/Q- and N-type) channels are important to ON activation of OB neurons (Isaacson and Strowbridge, 1998). Therefore, zinc released by OSNs in the GL may influence this activation. Calcium influx via NMDARs and voltage-gated calcium (P/Q- and N-type) channels close to granule cell-mitral cell dendrodendritic synapses is associated with transmitter (GABA) release (Isaacson and Strowbridge, 1998; Chen et al., 2000; Halabisky et al., 2000; Isaacson, 2001; Egger et al., 2003, 2005). Thus, zinc released from centrifugal fiber terminals contacting granule cells may inhibit voltage-gated calcium channels expressed by these neurons, affecting excitability and/or indirectly influencing transmitter release.

Zinc’s Effects on Voltage-Gated Sodium Channels

Voltage-gated sodium channels are composed of a pore-forming alpha subunit and two auxiliary beta subunits that modulate channel gating (Catterall, 2000; Catterall et al., 2005). The individual voltage-gated sodium channel subtypes are defined by the nine different alpha channel subunits (Nav1.1–Nav1.9; Deuis et al., 2017). Several types of these subunits (e.g., Nav1.1, Nav1.6, Nav 1.7) are heterogeneously expressed in the rat olfactory system (Lorincz and Nusser, 2008; Ahn et al., 2011).

Whereas most voltage-gated sodium channels are highly sensitive to blockade by tetrodotoxin (TTX), some (e.g., Nav1.5, Nav1.8 and Nav1.9) are much less sensitive to TTX (Goldin, 2001; Mathie et al., 2006). A comparison of the sensitivity of sodium channels to zinc revealed potent inhibition of TTX-resistant channels by zinc (IC50 of 50 μM) compared to TTX-sensitive channels (IC50 of 2 mM; Frelin et al., 1986). Whereas Nav1.5 channels are frequently characterized as “‘cardiac”’ sodium channels, some neurons also express these channels (Mathie et al., 2006).

In rat medial entorhinal cortical neurons, both TTX-sensitive and TTX-resistant sodium channels were found, with the latter being highly sensitive to zinc (IC50 of around 9 μM; White et al., 1993; Mathie et al., 2006). At a higher concentration (around 300 μM), zinc inhibited inward TTX-resistant sodium channels in dorsal root ganglion neurons, which chiefly express Nav1.8 and Nav1.9 channels (Kuo et al., 2004).

In our 2001 study, we also examined the effects of zinc on voltage-gated sodium channels in cultured rat OB neurons (Horning and Trombley, 2001). We hypothesized that our observed suppression of excitatory transmission by zinc (Trombley et al., 1998; Horning and Trombley, 2001) may be mediated, in part, by actions on voltage-gated sodium channels, which are largely responsible for the upstroke of an action potential. Sodium currents were evoked under voltage-clamp conditions, and 100 μM zinc significantly decreased the sodium current by approximately 20%.

In our experiments in the OB, application of 1 μM TTX blocked these voltage-gated sodium currents; this finding confirms that these currents, which were inhibited by 100 μM zinc, were mediated by typical TTX-sensitive sodium channels. It is plausible that the OB contains a mixture of TTX-sensitive and TTX-resistant sodium channels. Given the findings of previous investigators, it is possible that lower concentrations of synaptic zinc may modulate TTX-resistant sodium channels in the OB.

Delgado et al. (2006) subsequently investigated the effects of several metals including zinc on TTX-sensitive voltage-dependent inward sodium currents in toad OSNs (Aedo et al., 2007). At a very low concentration (0.1 μM), zinc increased the amplitude and accelerated the kinetics of activation and inactivation. At a much higher concentration (100 μM), zinc inhibited the inward sodium current by 33% (Delgado et al., 2006). These investigators also found that low concentrations (1–50 μM) of zinc increased the neuronal firing rate in a concentration-dependent manner, while higher concentrations (100–500 μM) of zinc decreased the firing rate (Aedo et al., 2007). These findings are consistent with our findings in the OB.

The observed effects of zinc on voltage-gated sodium channels have implications to OB function. When rat OB mitral cells are held near threshold, they can generate rebound spikes that require recovery of subthreshold TTX-sensitive sodium channels (Balu and Strowbridge, 2007). These investigators further showed that currents mediated by these channels contribute to rebound spikes following granule cell-mediated IPSPs by enhancing subthreshold excitatory events. Thus, zinc-mediated inhibition of these voltage-gated sodium channels could influence the bidirectional control of spike output from the OB.

ET cells have a TTX-sensitive persistent sodium current (Hayar et al., 2004b; Liu and Shipley, 2008b), which is necessary for spontaneous burst initiation in ET cells (Hayar et al., 2004b). Liu and Shipley (2008a) investigated the actions of this current on excitatory synaptic transmission from ON to ET cells in mouse OB. Collectively, their results suggest that this persistent sodium current amplifies subthreshold postsynaptic excitatory responses (EPSPs) and produces temporal summation. Dopamine neurons in the mouse OB also have a mostly TTX-sensitive persistent sodium current that contributes to autorhythmicity of these neurons (Pignatelli et al., 2005). Given the well-documented importance of both ET and DA cells in regulating glomerular activity, zinc-mediated inhibition of this current could alter several critical glomerular circuit functions underlying odor information processing.

Zinc’s Effects on Voltage-Gated Potassium Channels

Voltage-gated potassium channels are the largest family of voltage-gated ion channels, consisting of ~40 subunits that form multiple distinct channels with different functional properties (Gutman et al., 2003, 2005; Nusser, 2009). They influence neuronal excitability in numerous ways including being responsible for the repolarization phase of action potentials and regulating action potential frequency. The focus of this discussion is on two subtypes: delayed rectifier and A-type potassium channels.

A variety of delayed-rectifier and A-type potassium channel subunits (e.g., Kv1.1, Kv1.2, Kv1.3, Kv1.4, Kv3.1, Kv3.2, Kv4.2, and Kv4.3) are heterogeneously expressed in the OB (Veh et al., 1995; Fadool et al., 2004; Kollo et al., 2008; Lorincz and Nusser, 2008; Boda et al., 2012). Several voltage-gated potassium channel subunits (Kv2.1, Kv3.1b, Kv4.3) are found in a high percentage (70%–95%) of deep SA cells (Eyre et al., 2009). It has been shown that intracellular zinc and calcium released in response to oxidative stimuli play a role in the insertion of the Kv2.1 channel into the plasma membranes of cortical neurons, detected by significant enhancement of the delayed rectifier potassium currents after brief exposure to apoptogenic stimuli (McLaughlin et al., 2001; Pal et al., 2003, 2006; Redman et al., 2007, 2009; McCord and Aizenman, 2013). This current enhancement leads to a loss of intracellular potassium that regulates processes involved in neuronal apoptosis (McCord and Aizenman, 2013).

Various potassium channels, including A-type and delayed-rectifier channels, are present or are probably expressed on granule cell dendrites in the OB (Veh et al., 1995; Schoppa and Westbrook, 1999; Bywalez et al., 2015), where A-type potassium channels play an important role in granule cell signaling (Schoppa and Westbrook, 1999). Bywalez et al. (2015) recently concluded that delayed-rectifier currents likely play a role in local repolarization in granule cell dendritic spines. Other types of potassium channels have important functions in the OB; these include large-conductance calcium–dependent K currents (IBK) in ET cells that terminate spontaneous bursting (Liu and Shipley, 2008b) and small-conductance calcium-activated potassium channels (SK) activated either by calcium channels or NMDARs in mitral cell dendrites that influence action potential firing and dendrodendritic inhibition (Maher and Westbrook, 2005). Thus, effects of zinc on potassium channels in the OB would have important implications to OB function.

Zinc’s Effects on Transient and Steady-State Outward Potassium Currents

Zinc inhibition of A-type channels was first shown in cultured rat sympathetic neurons (Constanti and Smart, 1987). It was subsequently shown that zinc (1 μM–1 mM) markedly alters the voltage dependence of activation and inactivation for A-type channels in rat cerebellar granule cells (Bardoni and Belluzzi, 1994). Zinc-induced shifts in the steady-state inactivation and activation curves to more positive potentials increased the peak current amplitude near resting membrane potentials and decreased the peak current at hyperpolarized potentials (Bardoni and Belluzzi, 1994). In another study, zinc (2 μM–1000 μM) modulated the gating (slowed activation) of delayed-rectifier channels cloned from rat and human tissues in a concentration-dependent manner (Harrison et al., 1993).

In our 2001 study, we also examined the effects of zinc on voltage-gated potassium channels in cultured rat OB neurons (Horning and Trombley, 2001). First, outward currents consistent with the kinetics of a delayed rectifier-type current were evoked by clamping neurons at −50 mV and stepping to +20 mV for 100 ms. Zinc (100 μM) significantly inhibited the steady-state current amplitudes but potentiated the peak current amplitudes in most neurons examined. This latter effect was due to the activation of a transient A-type outward current that was not observed in the absence of zinc.

The transient A-type potassium current helps mediate the interspike interval and is mostly activated by a depolarizing event that follows hyperpolarization. We examined this current by holding neurons at a hyperpolarizing potential (−90 mV) and stepping them to +20 mV for 100 ms, which activated both the A-type and delayed rectifier potassium currents. Zinc (100 μM) significantly inhibited both the transient component (A-type) and the steady-state component (delayed rectifier-type) of the potassium currents evoked under these conditions.

Our results suggest that zinc’s effects on potassium channels are influenced not only by the type of potassium channel but also by the membrane voltage. While zinc inhibited the delayed rectifier-type outward current at all voltages examined, its effects on the transient A-type current were more complex. Zinc (100 μM) inhibited A-type currents when the membrane voltage was hyperpolarized but enhanced the current when the membrane was at or depolarized to a typical resting potential where most A-type channels are inactivated.

Consistent with our findings in the OB, 50 μM zinc reduced the amplitude of delayed-rectifier currents and increased the peak amplitude of A-type currents in rat hDBB neurons (Easaw et al., 1999). Our observed voltage-dependent effects of zinc on A-type channels in OB neurons are also similar to the above-described findings in cerebellar neurons (Bardoni and Belluzzi, 1994).

Another study examined the effects of zinc on A-type currents in PG cells in the rat OB. Voltage-clamp analyses of PG cells revealed two subpopulations based on the characteristics of their potassium conductances (Puopolo and Belluzzi, 1998). The first group of PG cells displayed delayed-rectifier and fast transient A-type currents of similar amplitude. The second group of PG cells had a large A-type current and a small or absent delayed-rectifier current. In this study, zinc (10–300 μM) affected A-type, but not delayed-rectifier, channels. When the inactivation of A-type channels was removed with a hyperpolarizing step to −120 mV, zinc caused a slight reduction in the current amplitude Similar to our results, the effect of zinc on the A-type current was potentiation when the membrane was depolarized starting from physiological holding potentials (e.g., −50 mV; Puopolo and Belluzzi, 1998).

Zinc’s Effects on Potassium Channels and Repetitive Firing

As the delayed rectifier-type current is involved in the repolarization phase of action potentials, its inhibition would likely reduce the rate of repetitive firing. Potentiation of the A-type outward current would slow the depolarization rate, thus, also reduce the repetitive firing rate. Consistent with this prediction, at a holding potential of −50 mV, 100 μM zinc reduced neuronal firing in all neurons examined by about 50% (Horning and Trombley, 2001). At a holding potential of −90 mV, 100 μM zinc decreased the number of action potentials fired by most (~70%) neurons by about 63%. However, zinc also increased action potential frequency in a significant proportion (~30%) of neurons at this holding potential. Thus, zinc appears to modulate repetitive firing in a voltage-dependent manner, with some variability perhaps due to the diversity of potassium channel expression among OB neurons.

The effects of zinc on potassium currents in the OB have implications to bulb function. There exists a delay in the onset of action potential firing in M/T cells, which may be attributable to A-type potassium currents (Chen and Shepherd, 1997). As zinc-containing OSNs synapse with mitral, tufted, and PG cells, zinc co-released with glutamate may affect glomerular circuit activity via effects on these A-type currents. In addition, zinc’s ability to dramatically alter the excitability of the subset of PG cells that express primarily the A-type channel (Puopolo and Belluzzi, 1998; Fogli Iseppe et al., 2016) could significantly alter glomerular circuit function via regulation of feedback and feedforward inhibition from these cells.

A-type currents are also responsible for a delay in action potential firing in granule cells, thus, influence the timing of reciprocal inhibition (Schoppa and Westbrook, 1999). As granule cells appear to be targets of zinc-containing centrifugal fibers, zinc’s opposing actions on A-type currents also may indirectly affect reciprocal inhibition mediated by these cells.

Potential Effects of Zinc Signaling on Behavior

Results from ZnT3 KO Mice

Some previous studies have examined the effects of application of zinc or zinc chelators on behavior. Intrathecal injection of zinc or zinc chelators (CaEDTA, dipicolinic acid) altered nociceptive and antinociceptive activity in mice (Larson and Kitto, 1997, 1999). Infusion of the zinc chelator diethyldithiocarbamate into the hippocampus disrupted spatial-working memory (Frederickson et al., 1990) and spatial learning (Lassalle et al., 2000) in rodents. In a later study, diethyldithiocarbamate or CaEDTA infusion into the hippocampus of mice disrupted the acquisition and consolidation of contextual fear conditioning, without affecting retrieval (Daumas et al., 2004). However, the development of the ZnT3 KO mouse has provided a novel tool to test the hypothesis that zinc signaling may modulate behavior, as reviewed by McAllister and Dyck (2017).

Early behavioral characterization of ZnT3 KO mice by Cole et al. (2001) revealed minimal abnormalities in a variety of behaviors including those involving motor coordination, thermal nociception, olfaction, auditory startle response, anxiety, and certain forms of learning and memory (e.g., passive avoidance, fear conditioning, spatial learning, working and reference memory; Cole et al., 2001; McAllister and Dyck, 2017). More recent evidence showed differences between ZnT3 KO mice and control mice in certain behavioral tests assessing fear conditioning and memory (Martel et al., 2010) and social memory (Martel et al., 2011). Differences between ZnT3 KO mice and control mice in performance on the standard water maze further suggest altered spatial navigation and memory in ZnT3 KO mice (Martel et al., 2011; McAllister and Dyck, 2017). A Sindreu et al. (2011) study found deficits in contextual discrimination and spatial working memory in ZnT3 KO mice. Perhaps the greatest evidence of behavioral abnormalities in ZnT3 KO mice comes from a recent study by Yoo et al. (2016) who observed sex-dependent behavioral abnormalities in young male ZnT3 KO mice consistent with autistic behavior, which included reduced social interaction, decreased novelty seeking behavior, increased repetitive behavior and reduced locomotion.

Little is known about the potential effects of zinc modulation in the OB on olfactory behavior. In the study by Cole et al. (2001), olfaction in ZnT3 KO mice was measured in two ways: (1) measurement of latency to find food hidden beneath bedding to assess gross olfactory deficits; and (2) a conditioning paradigm requiring mice to associate a non-noxious odor (peppermint) with an aversive compound (quinine) to assess the olfactory threshold (Cole et al., 2001). ZnT3 KO and control mice had similar food-recovery latencies and olfactory thresholds. However, they acknowledge there are some limitations to this study. For example, the tests employed may not be the most sensitive to neuromodulation by zinc. As they indicated, KO mice also may compensate during development by altering neuronal anatomy, neurotransmitter release, or receptor sensitivity (Cole et al., 2001). Thus, they suggest that zinc may be an effective neuromodulator in wild-type mice, but KO mice could compensate easily for its absence. Furthermore, some behavioral abnormalities are strongly sex-dependent (Yoo et al., 2016), and this variable was not examined in the olfactory studies by Cole et al. (2001).

Results from Progranulin-Deficient (KO) Mice

Progranulin is a protein with both neuroprotective and immunological properties (Rademakers et al., 2012), and its functions have been proposed to include protein homeostasis (Hardt et al., 2017). As progranulin deficiency in humans is linked to frontotemporal lobar degeneration (Baker et al., 2006; Cruts et al., 2006). Hardt et al. (2017) assessed psychopathology and chronic pain in progranulin-deficient (KO) mice. They present novel data suggesting that defective neuronal zinc transport may play a role in psychopathology associated with progranulin deficiency, which includes abnormalities in olfactory behavior.

In addition to assessing behaviors related to memory, emotions (e.g., anxiety, depression) and nociception, Hardt et al. (2017) assessed odor preference and discrimination in progranulin KO and control mice. Interest and preferences of odors prior to and following peripheral nerve injury were assessed using cinnamon (familiar odor) and vanilla (novel odor). Whereas control animals showed no loss of interest in pleasant odors after nerve injury, progranulin KO mice increased visit frequency of odors after nerve injury, leading to longer traveling paths. Odor discrimination was assessed using rose (novel odor) vs. vanilla (familiar odor). In contrast to controls, progranulin KO mice did not discriminate between these odors.

Due to the abnormal odor-related behavior in progranulin KO mice, these investigators performed deep proteome analyses of the prefrontal cortexes and OBs of progranulin KO and control mice (Hardt et al., 2017). This revealed progranulin-dependent changes in proteins involved in synaptic transport, with a loss of nuclear and synaptic zinc transporters (ZnT9/Slc30a9; ZnT3/Slc30a3) associated with progranulin deficiency. Western Blot analysis of the OB, prefrontal cortex and hippocampus revealed a complete loss of ZnT3/Slc30a3 in progranulin KO mice along with decreased plasma zinc levels (Hardt et al., 2017). Collectively, these results suggest that previously unrecognized zinc-dependent molecular mechanisms underlie progranulin-related loss of function, including deficits in odor preference and discrimination.

Concluding Remarks

While most attention has focused on zinc’s effects on neuronal excitability and synaptic transmission in other regions of the CNS, comparatively little is known about zinc’s effects in the OB (see Supplementary Tables S1, S2). This opens the door for many future directions of research into mechanisms by which zinc could modulate odor information processing.

For example, NMDARs have a multitude of important functions, ranging from regulating mitral/tufted cell spiking to dendrodendritic inhibition. GABAergic transmission is also critical to both intra- and inter-glomerular inhibition and synchronization. It is clear that zinc is an effective modulator of both of these receptor types, but both the circuit details and the broad implications of such modulation remain to be elucidated.

The impact of zinc on OB circuit behavior due to modulation of AMPARs and glycine receptors is much more complex. This is due to the exceptional diversity of zinc’s modulatory effects (potentiation and/or inhibition) on subpopulations of these receptors in the OB, which are influenced by factors including the receptor molecular composition, the extracellular environment, the (desensitized) state of the receptor, as well as intracellular and extracellular sites of zinc action.

Zinc’s effects on voltage-gated ion (e.g., calcium, sodium, potassium) channels add another significant dimension, as numerous subtypes of these channels have important regulatory function on OB neuron excitability and network function.

Substantial progress has been made in the initial characterization of what zinc can do to modify excitability and synaptic transmission. However, it simply just scratches the surface of what is needed to be known to fully appreciate the impact of zinc on odor processing by the OB. This also highlights a significant aspect of using the OB to gain a broader understanding of the role of zinc as a neuromodulator. While the bulk of what is known about zinc-mediated neuromodulation comes from other brain regions (e.g., hippocampus), the functions of those brain regions and the subcircuits contained within (e.g., MFs) are not entirely well defined. A major advantage of using the OB to investigate the actions of zinc is that we know precisely the function of the OSN → OB glomerular network (where most of the zinc is located), that is—the processing of odor information. We would be hard pressed to identify a brain region more lucrative for identifying mechanisms involved in modifying/regulating information processing. This, in itself, makes the OB a unique tool for understanding the significance of zinc-mediated neuromodulation in the CNS.

Author Contributions

Both authors contributed equally to this manuscript. LJB wrote the first draft of the manuscript and PQT edited this draft and contributed to all subsequent drafts. All authors agreed upon the final version of this manuscript.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

AMPA, alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; AMPARs, AMPA receptors; CNS, central nervous system; EPL, external plexiform layer (olfactory bulb); EPSC, excitatory postsynaptic current; EPSP, excitatory postsynaptic potential; ET cells, external tufted cells; GABA, gamma-aminobutyric acid; GCL, granule cell layer (olfactory bulb); GL, glomerular layer (olfactory bulb); hDBB, horizontal limb of the diagonal band of Broca; HVA, high-voltage activated; IC50, half maximal inhibitory concentration; IPL, internal plexiform layer (olfactory bulb); IPSP, inhibitory postsynaptic potential; JG cells, juxtaglomerular cells; KO, knockout; LTP, long-term potentiation; LVA, low-voltage activated; MCL, mitral cell layer (olfactory bulb); MF, mossy fiber (hippocampus); M/T cells, mitral/tufted cells; NMDA, N-methyl-D aspartate; NMDARs, NMDA receptors; OB, olfactory bulb; ONL, olfactory nerve layer (olfactory bulb); OSN, olfactory sensory neuron; PG cells, periglomerular cells; SA cells, short axon cells; TPA, tris(2-pyridylmethyl)amine; TPEN, N,N,N′,N′-tetrakis (2-pyridylmethyl)ethylenediamine; TTX, tetrodotoxin.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/article/10.3389/fncel.2017.00297/full#supplementary-material

References

Aedo, F., Delgado, R., Wolff, D., and Vergara, C. (2007). Copper and zinc as modulators of neuronal excitability in a physiologically significant concentration range. Neurochem. Int. 50, 591–600. doi: 10.1016/j.neuint.2006.12.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Ahn, H. S., Black, J. A., Zhao, P., Tyrrell, L., Waxman, S. G., and Dib-Hajj, S. D. (2011). Nav1.7 is the predominant sodium channel in rodent olfactory sensory neurons. Mol. Pain 7:32. doi: 10.1186/1744-8069-7-32

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, C. T., Radford, R. J., Zastrow, M. L., Zhang, D. Y., Apfel, U. P., Lippard, S. J., et al. (2015). Modulation of extrasynaptic NMDA receptors by synaptic and tonic zinc. Proc. Natl. Acad. Sci. U S A 112, E2705–E2714. doi: 10.1073/pnas.1503348112

PubMed Abstract | CrossRef Full Text | Google Scholar

Aroniadou-Anderjaska, V., Ennis, M., and Shipley, M. T. (1997). Glomerular synaptic responses to olfactory nerve input in rat olfactory bulb slices. Neuroscience 79, 425–434. doi: 10.1016/s0306-4522(96)00706-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Aroniadou-Anderjaska, V., Ennis, M., and Shipley, M. T. (1999). Dendrodendritic recurrent excitation in mitral cells of the rat olfactory bulb. J. Neurophysiol. 82, 489–494.

PubMed Abstract | Google Scholar

Assaf, S. Y., and Chung, S. H. (1984). Release of endogenous Zn2+ from brain tissue during activity. Nature 308, 734–736. doi: 10.1038/308734a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Aungst, J. L., Heyward, P. M., Puche, A. C., Karnup, S. V., Hayar, A., Szabo, G., et al. (2003). Centre-surround inhibition among olfactory bulb glomeruli. Nature 426, 623–629. doi: 10.1038/nature02185

PubMed Abstract | CrossRef Full Text | Google Scholar

Baker, M., Mackenzie, I. R., Pickering-Brown, S. M., Gass, J., Rademakers, R., Lindholm, C., et al. (2006). Mutations in progranulin cause tau-negative frontotemporal dementia linked to chromosome 17. Nature 442, 916–919. doi: 10.1038/nature05016

PubMed Abstract | CrossRef Full Text | Google Scholar

Balu, R., and Strowbridge, B. W. (2007). Opposing inward and outward conductances regulate rebound discharges in olfactory mitral cells. J. Neurophysiol. 97, 1959–1968. doi: 10.1152/jn.01115.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Bardoni, R., and Belluzzi, O. (1994). Modifications of A-current kinetics in mammalian central neurones induced by extracellular zinc. J. Physiol. 479, 389–400. doi: 10.1113/jphysiol.1994.sp020304

PubMed Abstract | CrossRef Full Text | Google Scholar

Barnard, E. A., Darlison, M. G., Fujita, N., Glencorse, T. A., Levitan, E. S., Reale, V., et al. (1988). Molecular biology of the GABAA receptor. Adv. Exp. Med. Biol. 236, 31–45. doi: 10.1007/978-1-4757-5971-6_3

PubMed Abstract | CrossRef Full Text | Google Scholar

Beato, M. (2008). The time course of transmitter at glycinergic synapses onto motoneurons. J. Neurosci. 28, 7412–7425. doi: 10.1523/JNEUROSCI.0581-08.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Beaulieu, C., Dyck, R., and Cynader, M. (1992). Enrichment of glutamate in zinc-containing terminals of the cat visual cortex. Neuroreport 3, 861–864. doi: 10.1097/00001756-199210000-00010

PubMed Abstract | CrossRef Full Text

Berkowicz, D. A., and Trombley, P. Q. (2000). Dopaminergic modulation at the olfactory nerve synapse. Brain Res. 855, 90–99. doi: 10.1016/s0006-8993(99)02342-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Berkowicz, D. A., Trombley, P. Q., and Shepherd, G. M. (1994). Evidence for glutamate as the olfactory receptor cell neurotransmitter. J. Neurophysiol. 71, 2557–2561.

PubMed Abstract | Google Scholar

Birinyi, A., Parker, D., Antal, M., and Shupliakov, O. (2001). Zinc co-localizes with GABA and glycine in synapses in the lamprey spinal cord. J. Comp. Neurol. 433, 208–221. doi: 10.1002/cne.1136

PubMed Abstract | CrossRef Full Text | Google Scholar

Bitanihirwe, B. K., and Cunningham, M. G. (2009). Zinc: the brain’s dark horse. Synapse 63, 1029–1049. doi: 10.1002/syn.20683

PubMed Abstract | CrossRef Full Text | Google Scholar

Blakemore, L. J., Resasco, M., Mercado, M. A., and Trombley, P. Q. (2006). Evidence for Ca2+-permeable AMPA receptors in the olfactory bulb. Am. J. Physiol. Cell Physiol. 290, C925–C935. doi: 10.1152/ajpcell.00392.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Blakemore, L. J., Tomat, E., Lippard, S. J., and Trombley, P. Q. (2013). Zinc released from olfactory bulb glomeruli by patterned electrical stimulation of the olfactory nerve. Metallomics 5, 208–213. doi: 10.1039/c3mt20158a

PubMed Abstract | CrossRef Full Text | Google Scholar

Blakemore, L. J., and Trombley, P. Q. (2003). Kinetic variability of AMPA receptors among olfactory bulb neurons in culture. Neuroreport 14, 965–970. doi: 10.1097/01.wnr.0000070826.57864.00

PubMed Abstract | CrossRef Full Text | Google Scholar

Blakemore, L. J., and Trombley, P. Q. (2004). Diverse modulation of olfactory bulb AMPA receptors by zinc. Neuroreport 15, 919–923. doi: 10.1097/00001756-200404090-00037

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloomenthal, A. B., Goldwater, E., Pritchett, D. B., and Harrison, N. L. (1994). Biphasic modulation of the strychnine-sensitive glycine receptor by Zn2+. Mol. Pharmacol. 46, 1156–1159.

PubMed Abstract | Google Scholar

Boda, E., Hoxha, E., Pini, A., Montarolo, F., and Tempia, F. (2012). Brain expression of Kv3 subunits during development, adulthood and aging and in a murine model of Alzheimer’s disease. J. Mol. Neurosci. 46, 606–615. doi: 10.1007/s12031-011-9648-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Bresink, I., Ebert, B., Parsons, C. G., and Mutschler, E. (1996). Zinc changes AMPA receptor properties: results of binding studies and patch clamp recordings. Neuropharmacology 35, 503–509. doi: 10.1016/0028-3908(95)00192-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Burdette, S. C., Walkup, G. K., Spingler, B., Tsien, R. Y., and Lippard, S. J. (2001). Fluorescent sensors for Zn2+ based on a fluorescein platform: synthesis, properties and intracellular distribution. J. Am. Chem. Soc. 123, 7831–7841. doi: 10.1021/ja010059l

PubMed Abstract | CrossRef Full Text | Google Scholar

Burnashev, N., Zhou, Z., Neher, E., and Sakmann, B. (1995). Fractional calcium currents through recombinant GluR channels of the NMDA, AMPA and kainate receptor subtypes. J. Physiol. 485, 403–418. doi: 10.1113/jphysiol.1995.sp020738

PubMed Abstract | CrossRef Full Text | Google Scholar

Büsselberg, D., Michael, D., Evans, M. L., Carpenter, D. O., and Haas, H. L. (1992). Zinc (Zn2+) blocks voltage gated calcium channels in cultured rat dorsal root ganglion cells. Brain Res. 593, 77–81. doi: 10.1016/0006-8993(92)91266-h

PubMed Abstract | CrossRef Full Text | Google Scholar

Büsselberg, D., Platt, B., Michael, D., Carpenter, D. O., and Haas, H. L. (1994). Mammalian voltage-activated calcium channel currents are blocked by Pb2+, Zn2+, and Al3+. J. Neurophysiol. 71, 1491–1497.

PubMed Abstract | Google Scholar

Bywalez, W. G., Patirniche, D., Rupprecht, V., Stemmler, M., Herz, A. V., Pálfi, D., et al. (2015). Local postsynaptic voltage-gated sodium channel activation in dendritic spines of olfactory bulb granule cells. Neuron 85, 590–601. doi: 10.1016/j.neuron.2014.12.051

PubMed Abstract | CrossRef Full Text | Google Scholar

Carlson, G. C., Shipley, M. T., and Keller, A. (2000). Long-lasting depolarizations in mitral cells of the rat olfactory bulb. J. Neurosci. 20, 2011–2021.

PubMed Abstract | Google Scholar

Catterall, W. A. (2000). From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron 26, 13–25. doi: 10.1016/S0896-6273(00)81133-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Catterall, W. A., Goldin, A. L., and Waxman, S. G. (2005). International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol. Rev. 57, 397–409. doi: 10.1124/pr.57.4.4

PubMed Abstract | CrossRef Full Text

Chang, C. J., and Lippard, S. J. (2006). “Metalloneurochemistry: physiology, pathology, and probes,” in Neurodegenerative Diseases and Metal Ions, eds A. Sigel, H. Sigel and R. K. O. Sigel (Chichester, England: John Wiley and Son, Ltd), 321–370.

Google Scholar

Chávez, A. E., Singer, J. H., and Diamond, J. S. (2006). Fast neurotransmitter release triggered by Ca influx through AMPA-type glutamate receptors. Nature 443, 705–708. doi: 10.1038/nature05123

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, N., Moshaver, A., and Raymond, L. A. (1997). Differential sensitivity of recombinant N-methyl-D-aspartate receptor subtypes to zinc inhibition. Mol. Pharmacol. 51, 1015–1023.

PubMed Abstract | Google Scholar

Chen, W. R., and Shepherd, G. M. (1997). Membrane and synaptic properties of mitral cells in slices of rat olfactory bulb. Brain Res. 745, 189–196. doi: 10.1016/s0006-8993(96)01150-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, W. R., Xiong, W., and Shepherd, G. M. (2000). Analysis of relations between NMDA receptors and GABA release at olfactory bulb reciprocal synapses. Neuron 25, 625–633. doi: 10.1016/s0896-6273(00)81065-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Choi, D. W., and Koh, J. Y. (1998). Zinc and brain injury. Annu. Rev. Neurosci. 21, 347–375. doi: 10.1146/annurev.neuro.21.1.347

PubMed Abstract | CrossRef Full Text | Google Scholar

Christie, J. M., Schoppa, N. E., and Westbrook, G. L. (2001). Tufted cell dendrodendritic inhibition in the olfactory bulb is dependent on NMDA receptor activity. J. Neurophysiol. 85, 169–173.

PubMed Abstract | Google Scholar

Christie, J. M., and Westbrook, G. L. (2006). Lateral excitation within the olfactory bulb. J. Neurosci. 26, 2269–2277. doi: 10.1523/JNEUROSCI.4791-05.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Clements, J. D., Lester, R. A., Tong, G., Jahr, C. E., and Westbrook, G. L. (1992). The time course of glutamate in the synaptic cleft. Science 258, 1498–1501. doi: 10.1126/science.1359647

PubMed Abstract | CrossRef Full Text | Google Scholar

Cole, T. B., Martyanova, A., and Palmiter, R. D. (2001). Removing zinc from synaptic vesicles does not impair spatial learning, memory, or sensorimotor functions in the mouse. Brain Res. 891, 253–265. doi: 10.1016/s0006-8993(00)03220-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Cole, T. B., Wenzel, H. J., Kafer, K. E., Schwartzkroin, P. A., and Palmiter, R. D. (1999). Elimination of zinc from synaptic vesicles in the intact mouse brain by disruption of the ZnT3 gene. Proc. Natl. Acad. Sci. U S A 96, 1716–1721. doi: 10.1073/pnas.96.4.1716

PubMed Abstract | CrossRef Full Text | Google Scholar

Collingridge, G. L., Olsen, R. W., Peters, J., and Spedding, M. (2009). A nomenclature for ligand-gated ion channels. Neuropharmacology 56, 2–5. doi: 10.1016/j.neuropharm.2008.06.063

PubMed Abstract | CrossRef Full Text | Google Scholar

Constanti, A., and Smart, T. G. (1987). Zinc blocks the A-current in cultured rat sympathetic neurones. J. Physiol. 396:159P.

Google Scholar

Corthell, J. T., Stathopoulos, A. M., Watson, C. C., Bertram, R., and Trombley, P. Q. (2013). Olfactory bulb monoamine concentrations vary with time of day. Neuroscience 247, 234–241. doi: 10.1016/j.neuroscience.2013.05.040

PubMed Abstract | CrossRef Full Text | Google Scholar

Crawford, I. L., and Connor, J. D. (1973). Localization and release of glutamic acid in relation to the hippocampal mossy fibre pathway. Nature 244, 442–443. doi: 10.1038/244442a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Cruts, M., Gijselinck, I., van der Zee, J., Engelborghs, S., Wils, H., Pirici, D., et al. (2006). Null mutations in progranulin cause ubiquitin-positive frontotemporal dementia linked to chromosome 17q21. Nature 442, 920–924. doi: 10.1038/nature05017

PubMed Abstract | CrossRef Full Text | Google Scholar

Csermely, P., Szamel, M., Resch, K., and Somogyi, J. (1988). Zinc can increase the activity of protein kinase C and contributes to its binding to plasma membranes in T lymphocytes. J. Biol. Chem. 263, 6487–6490.

PubMed Abstract | Google Scholar

Danscher, G. (1981). Histochemical demonstration of heavy metals. A revised version of the sulphide silver method suitable for both light and electronmicroscopy. Histochemistry 71, 1–16. doi: 10.1007/bf00592566

PubMed Abstract | CrossRef Full Text | Google Scholar

Danscher, G. (1996). The autometallographic zinc-sulphide method. A new approach involving in vivo creation of nanometer-sized zinc sulphide crystal lattices in zinc-enriched synaptic and secretory vesicles. Histochem. J. 28, 361–373. doi: 10.1007/bf02331399

PubMed Abstract | CrossRef Full Text | Google Scholar

Danscher, G., Howell, G., Pérez-Clausell, J., and Hertel, N. (1985). The dithizone, Timm’s sulphide silver and the selenium methods demonstrate a chelatable pool of zinc in CNS. A proton activation (PIXE) analysis of carbon tetrachloride extracts from rat brains and spinal cords intravitally treated with dithizone. Histochemistry 83, 419–422. doi: 10.1007/bf00509203

PubMed Abstract | CrossRef Full Text | Google Scholar

Darcy, D. P., and Isaacson, J. S. (2010). Calcium-permeable AMPA receptors mediate glutamatergic signaling in neural precursor cells of the postnatal olfactory bulb. J. Neurophysiol. 103, 1431–1437. doi: 10.1152/jn.00821.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Daumas, S., Halley, H., and Lassalle, J. M. (2004). Disruption of hippocampal CA3 network: effects on episodic-like memory processing in C57BL/6J mice. Eur. J. Neurosci. 20, 597–600. doi: 10.1111/j.1460-9568.2004.03484.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Davila, N. G., Blakemore, L. J., and Trombley, P. Q. (2003). Dopamine modulates synaptic transmission between rat olfactory bulb neurons in culture. J. Neurophysiol. 90, 395–404. doi: 10.1152/jn.01058.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Delgado, R., Vergara, C., and Wolff, D. (2006). Divalent cations as modulators of neuronal excitability: emphasis on copper and zinc. Biol. Res. 39, 173–182. doi: 10.4067/s0716-97602006000100019

PubMed Abstract | CrossRef Full Text | Google Scholar

de Olmos, J., Hardy, H., and Heimer, L. (1978). The afferent connections of the main and the accessory olfactory bulb formations in the rat: an experimental HRP-study. J. Comp. Neurol. 181, 213–244. doi: 10.1002/cne.901810202

PubMed Abstract | CrossRef Full Text | Google Scholar

De Saint Jan, D., Hirnet, D., Westbrook, G. L., and Charpak, S. (2009). External tufted cells drive the output of olfactory bulb glomeruli. J. Neurosci. 29, 2043–2052. doi: 10.1523/JNEUROSCI.5317-08.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Deuis, J. R., Mueller, A., Israel, M. R., and Vetter, I. (2017). The pharmacology of voltage-gated sodium channel activators. Neuropharmacology doi: 10.1016/j.neuropharm.2017.04.014 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Dietz, S. B., and Murthy, V. N. (2005). Contrasting short-term plasticity at two sides of the mitral-granule reciprocal synapse in the mammalian olfactory bulb. J. Physiol. 569, 475–488. doi: 10.1113/jphysiol.2005.095844

PubMed Abstract | CrossRef Full Text | Google Scholar

Dingledine, R., Borges, K., Bowie, D., and Traynelis, S. F. (1999). The glutamate receptor ion channels. Pharmacol. Rev. 51, 7–61.

PubMed Abstract | Google Scholar

Donaldson, J., Pierre, T. S., Minnich, J. L., and Barbeau, A. (1973). Determination of Na+, K+, Mg2+, Cu2+, Zn2+, and Mn2+ in rat brain regions. Can. J. Biochem. 51, 87–92. doi: 10.1139/o73-010

PubMed Abstract | CrossRef Full Text | Google Scholar

Dreixler, J. C., and Leonard, J. P. (1994). Subunit-specific enhancement of glutamate receptor responses by zinc. Mol. Brain Res. 22, 144–150. doi: 10.1016/0169-328x(94)90042-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Dreixler, J. C., and Leonard, J. P. (1997). Effects of external calcium on zinc modulation of AMPA receptors. Brain Res. 752, 170–174. doi: 10.1016/s0006-8993(96)01439-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Dyck, R., Beaulieu, C., and Cynader, M. (1993). Histochemical localization of synaptic zinc in the developing cat visual cortex. J. Comp. Neurol. 329, 53–67. doi: 10.1002/cne.903290105

PubMed Abstract | CrossRef Full Text | Google Scholar

Easaw, J. C., Jassar, B. S., Harris, K. H., and Jhamandas, J. H. (1999). Zinc modulation of ionic currents in the horizontal limb of the diagonal band of Broca. Neuroscience 94, 785–795. doi: 10.1016/s0306-4522(99)00308-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Egger, V., Svoboda, K., and Mainen, Z. F. (2003). Mechanisms of lateral inhibition in the olfactory bulb: efficiency and modulation of spike-evoked calcium influx into granule cells. J. Neurosci. 23, 7551–7558.

PubMed Abstract | Google Scholar

Egger, V., Svoboda, K., and Mainen, Z. F. (2005). Dendrodendritic synaptic signals in olfactory bulb granule cells: local spine boost and global low-threshold spike. J. Neurosci. 25, 3521–3530. doi: 10.1523/JNEUROSCI.4746-04.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Ennis, M., Hamilton, K. A., and Hayar, A. (2007). “Neurochemistry of the main olfactory system,” in Handbook of Neurochemistry and Molecular Neurobiology: Sensory Neurochemistry, 3rd Edn. ed. D. A. Johnson (Heidelberg, Germany: Springer), 137–204.

Google Scholar

Ennis, M., Puche, A. C., Holy, T., and Shipley, M. T. (2015). “The olfactory system,” in The Rat Nervous System: Fourth Edition, ed. G. Paxinos (London, UK: Elsevier Inc.), 761–803.

Google Scholar

Ennis, M., Zhou, F. M., Ciombor, K. J., Aroniadou-Anderjaska, V., Hayar, A., Borrelli, E., et al. (2001). Dopamine D2 receptor-mediated presynaptic inhibition of olfactory nerve terminals. J. Neurophysiol. 86, 2986–2997. doi: 10.1016/s0028-3908(01)00038-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Ennis, M., Zimmer, L. A., and Shipley, M. T. (1996). Olfactory nerve stimulation activates rat mitral cells via NMDA and non-NMDA receptors in vitro. Neuroreport 7, 989–992. doi: 10.1097/00001756-199604100-00007

PubMed Abstract | CrossRef Full Text | Google Scholar

Eyre, M. D., Antal, M., and Nusser, Z. (2008). Distinct deep short-axon cell subtypes of the main olfactory bulb provide novel intrabulbar and extrabulbar GABAergic connections. J. Neurosci. 28, 8217–8229. doi: 10.1523/JNEUROSCI.2490-08.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Eyre, M. D., Kerti, K., and Nusser, Z. (2009). Molecular diversity of deep short-axon cells of the rat main olfactory bulb. Eur. J. Neurosci. 29, 1397–1407. doi: 10.1111/j.1460-9568.2009.06703.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Fadool, D. A., Tucker, K., Perkins, R., Fasciani, G., Thompson, R. N., Parsons, A. D., et al. (2004). Kv1.3 channel gene-targeted deletion produces “Super-Smeller Mice” with altered glomeruli, interacting scaffolding proteins, and biophysics. Neuron 41, 389–404. doi: 10.1016/s0896-6273(03)00844-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Fekete, A., Johnston, J., and Delaney, K. R. (2014). Presynaptic T-type Ca2+ channels modulate dendrodendritic mitral-mitral and mitral-periglomerular connections in mouse olfactory bulb. J. Neurosci. 34, 14032–14045. doi: 10.1523/JNEUROSCI.0905-14.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Fogli Iseppe, A., Pignatelli, A., and Belluzzi, O. (2016). Calretinin-periglomerular interneurons in mice olfactory bulb: cells of few words. Front. Cell. Neurosci. 10:231. doi: 10.3389/fncel.2016.00231

PubMed Abstract | CrossRef Full Text | Google Scholar

Foster, T. C. (2007). Calcium homeostasis and modulation of synaptic plasticity in the aged brain. Aging Cell 6, 319–325. doi: 10.1111/j.1474-9726.2007.00283.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, C. J. (1989). Neurobiology of zinc and zinc-containing neurons. Int. Rev. Neurobiol. 31, 145–238. doi: 10.1016/s0074-7742(08)60279-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, R. E., Frederickson, C. J., and Danscher, G. (1990). In situ binding of bouton zinc reversibly disrupts performance on a spatial memory task. Behav. Brain Res. 38, 25–33. doi: 10.1016/0166-4328(90)90021-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, C. J., Klitenick, M. A., Manton, W. I., and Kirkpatrick, J. B. (1983). Cytoarchitectonic distribution of zinc in the hippocampus of man and the rat. Brain Res. 273, 335–339. doi: 10.1016/0006-8993(83)90858-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, C. J., Koh, J. Y., and Bush, A. I. (2005). The neurobiology of zinc in health and disease. Nat. Rev. Neurosci. 6, 449–462. doi: 10.1038/nrn1671

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, C. J., and Moncrieff, D. W. (1994). Zinc-containing neurons. Biol. Signals 3, 127–139. doi: 10.1159/000109536

PubMed Abstract | CrossRef Full Text | Google Scholar

Frederickson, C. J., Suh, S. W., Silva, D., Frederickson, C. J., and Thompson, R. B. (2000). Importance of zinc in the central nervous system: the zinc-containing neuron. J. Nutr. 130, 1471S–1483S.

PubMed Abstract | Google Scholar

Frelin, C., Cognard, C., Vigne, P., and Lazdunski, M. (1986). Tetrodotoxin-sensitive and tetrodotoxin-resistant Na+ channels differ in their sensitivity to Cd2+ and Zn2+. Eur. J. Pharmacol. 122, 245–250. doi: 10.1016/0014-2999(86)90109-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Friedman, B., and Price, J. L. (1984). Fiber systems in the olfactory bulb and cortex: a study in adult and developing rats, using the timm method with the light and electron microscope. J. Comp. Neurol. 223, 88–109. doi: 10.1002/cne.902230108

PubMed Abstract | CrossRef Full Text | Google Scholar

Fukada, T., Hojyo, S., and Bin, B. H. (2014). “Zinc signal in growth control and bone diseases,” in Zinc Signals in Cellular Functions and Disorders, eds T. Fukada and T. Kambe (Tokyo, Japan: Springer), 249–267.

Google Scholar

Gee, K. R., Zhou, Z. L., Qian, W. J., and Kennedy, R. (2002). Detection and imaging of zinc secretion from pancreatic β-cells using a new fluorescent zinc indicator. J. Am. Chem. Soc. 124, 776–778. doi: 10.1021/ja011774y

PubMed Abstract | CrossRef Full Text | Google Scholar

Ghosh, S. K., Kim, P., Zhang, X. A., Yun, S. H., Moore, A., Lippard, S. J., et al. (2010). A novel imaging approach for early detection of prostate cancer based on endogenous zinc sensing. Cancer Res. 70, 6119–6127. doi: 10.1158/0008-5472.CAN-10-1008

PubMed Abstract | CrossRef Full Text | Google Scholar

Gire, D. H., Franks, K. M., Zak, J. D., Tanaka, K. F., Whitesell, J. D., Mulligan, A. A., et al. (2012). Mitral cells in the olfactory bulb are mainly excited through a multistep signaling path. J. Neurosci. 32, 2964–2975. doi: 10.1523/JNEUROSCI.5580-11.2012

PubMed Abstract | CrossRef Full Text | Google Scholar

Giustetto, M., Bovolin, P., Fasolo, A., Bonino, M., Cantino, D., and Sassoe-Pognetto, M. (1997). Glutamate receptors in the olfactory bulb synaptic circuitry: heterogeneity and synaptic localization of N-methyl-D-aspartate receptor subunit 1 and AMPA receptor subunit 1. Neuroscience 76, 787–798. doi: 10.1016/s0306-4522(96)00285-0

PubMed Abstract | CrossRef Full Text

Goldberg, J. M., Loas, A., and Lippard, S. J. (2016). Metalloneurochemistry and the pierian spring: ‘shallow draughts intoxicate the brain’. Isr. J. Chem. 56, 791–802. doi: 10.1002/ijch.201600034

PubMed Abstract | CrossRef Full Text | Google Scholar

Goldin, A. L. (2001). Resurgence of sodium channel research. Annu. Rev. Physiol. 63, 871–894. doi: 10.1146/annurev.physiol.63.1.871

PubMed Abstract | CrossRef Full Text | Google Scholar

Golgi, C. (1875). Sulla fina struttura dei bulbi olfactorii (On the fine structure of the olfactory bulb). Riv. Sper. Freniatr. Med. Leg. 1, 404–425.

Google Scholar

Gulya, K., Kovács, G. L., and Kása, P. (1991). Partial depletion of endogenous zinc level by (D-Pen2, D-Pen5)enkephalin in the rat brain. Life Sci. 48, PL57–PL62. doi: 10.1016/0024-3205(91)90462-k

PubMed Abstract | CrossRef Full Text | Google Scholar

Gutman, G. A., Chandy, K. G., Adelman, J. P., Aiyar, J., Bayliss, D. A., Clapham, D. E., et al. (2003). International Union of Pharmacology. XLI. Compendium of voltage-gated ion channels: potassium channels. Pharmacol. Rev. 55, 583–586. doi: 10.1124/pr.55.4.9

PubMed Abstract | CrossRef Full Text | Google Scholar

Gutman, G. A., Chandy, K. G., Grissmer, S., Lazdunski, M., McKinnon, D., Pardo, L. A., et al. (2005). International Union of Pharmacology. LIII. Nomenclature and molecular relationships of voltage-gated potassium channels. Pharmacol. Rev. 57, 473–508. doi: 10.1124/pr.57.4.10

PubMed Abstract | CrossRef Full Text | Google Scholar

Halabisky, B., Friedman, D., Radojicic, M., and Strowbridge, B. W. (2000). Calcium influx through NMDA receptors directly evokes GABA release in olfactory bulb granule cells. J. Neurosci. 20, 5124–5134.

PubMed Abstract | Google Scholar

Hara, T., Takeda, T. A., Takagishi, T., Fukue, K., Kambe, T., and Fukada, T. (2017). Physiological roles of zinc transporters: molecular and genetic importance in zinc homeostasis. J. Physiol. Sci. 67, 283–301. doi: 10.1007/s12576-017-0521-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Hardt, S., Heidler, J., Albuquerque, B., Valek, L., Altmann, C., Wilken-Schmitz, A., et al. (2017). Loss of synaptic zinc transport in progranulin deficient mice may contribute to progranulin-associated psychopathology and chronic pain. Biochim. Biophys. Acta doi: 10.1016/j.bbadis.2017.07.014 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Harrison, N. L., Radke, H. K., Tamkun, M. M., and Lovinger, D. M. (1993). Modulation of gating of cloned rat and human K+ channels by micromolar Zn2+. Mol. Pharmacol. 43, 482–486.

PubMed Abstract | Google Scholar

Harvey, R. J., Thomas, P., James, C. H., Wilderspin, A., and Smart, T. G. (1999). Identification of an inhibitory Zn2+ binding site on the human glycine receptor alpha1 subunit. J. Physiol. 520, 53–64. doi: 10.1111/j.1469-7793.1999.00053.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Haug, F. M. (1967). Electron microscopical localization of the zinc in hippocampal mossy fibre synapses by a modified sulfide silver procedure. Histochemie 8, 355–368. doi: 10.1007/bf00401978

PubMed Abstract | CrossRef Full Text | Google Scholar

Hayar, A., Karnup, S., Ennis, M., and Shipley, M. T. (2004a). External tufted cells: a major excitatory element that coordinates glomerular activity. J. Neurosci. 24, 6676–6685. doi: 10.1523/jneurosci.1367-04.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Hayar, A., Karnup, S., Shipley, M. T., and Ennis, M. (2004b). Olfactory bulb glomeruli: external tufted cells intrinsically burst at theta frequency and are entrained by patterned olfactory input. J. Neurosci. 24, 1190–1199. doi: 10.1523/jneurosci.4714-03.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Hayar, A., Shipley, M. T., and Ennis, M. (2005). Olfactory bulb external tufted cells are synchronized by multiple intraglomerular mechanisms. J. Neurosci. 25, 8197–8208. doi: 10.1523/jneurosci.2374-05.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Hirata, Y. (1964). Some observations on the fine structure of the synapses in the olfactory bulb of the mouse, with particular reference to the atypical synaptic configurations. Arch. Histol. Jpn. 24, 293–302. doi: 10.1679/aohc1950.24.293

PubMed Abstract | CrossRef Full Text | Google Scholar

Hirzel, K., Müller, U., Latal, A. T., Hülsmann, S., Grudzinska, J., Seeliger, M. W., et al. (2006). Hyperekplexia phenotype of glycine receptor alpha1 subunit mutant mice identifies Zn2+ as an essential endogenous modulator of glycinergic neurotransmission. Neuron 52, 679–690. doi: 10.1016/j.neuron.2006.09.035

PubMed Abstract | CrossRef Full Text | Google Scholar

Hojyo, S., and Fukada, T. (2016). Zinc transporters and signaling in physiology and pathogenesis. Arch. Biochem. Biophys. 611, 43–50. doi: 10.1016/j.abb.2016.06.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Horning, M. S., Kwon, B., Blakemore, L. J., Spencer, C. M., Goltz, M., Houpt, T. A., et al. (2004). Alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate receptor subunit expression in rat olfactory bulb. Neurosci. Lett. 372, 230–234. doi: 10.1016/j.neulet.2004.09.044

PubMed Abstract | CrossRef Full Text | Google Scholar

Horning, M. S., and Trombley, P. Q. (2001). Zinc and copper influence excitability of rat olfactory bulb neurons by multiple mechanisms. J. Neurophysiol. 86, 1652–1660.

PubMed Abstract | Google Scholar

Howell, G. A., Welch, M. G., and Frederickson, C. J. (1984). Stimulation-induced uptake and release of zinc in hippocampal slices. Nature 308, 736–738. doi: 10.1038/308736a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Hsia, A. Y., Vincent, J. D., and Lledo, P. M. (1999). Dopamine depresses synaptic inputs into the olfactory bulb. J. Neurophysiol. 82, 1082–1085.

PubMed Abstract | Google Scholar

Huang, Z., Zhang, X. A., Bosch, M., Smith, S. J., and Lippard, S. J. (2013). Tris(2-pyridylmethyl)amine (TPA) as a membrane-permeable chelator for interception of biological mobile zinc. Metallomics 5, 648–655. doi: 10.1039/c3mt00103b

PubMed Abstract | CrossRef Full Text | Google Scholar

Hume, R. I., Dingledine, R., and Heinemann, S. F. (1991). Identification of a site in glutamate receptor subunits that controls calcium permeability. Science 253, 1028–1031. doi: 10.1126/science.1653450

PubMed Abstract | CrossRef Full Text | Google Scholar

Ichinohe, N., and Rockland, K. S. (2005). Distribution of synaptic zinc in the macaque monkey amygdala. J. Comp. Neurol. 489, 135–147. doi: 10.1002/cne.20632

PubMed Abstract | CrossRef Full Text | Google Scholar

Isaacson, J. S. (2001). Mechanisms governing dendritic γ-aminobutyric acid (GABA) release in the rat olfactory bulb. Proc. Natl. Acad. Sci. U S A 98, 337–342. doi: 10.1073/pnas.021445798

PubMed Abstract | CrossRef Full Text | Google Scholar

Isaacson, J. S., and Strowbridge, B. W. (1998). Olfactory reciprocal synapses: dendritic signaling in the CNS. Neuron 20, 749–761. doi: 10.1016/s0896-6273(00)81013-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Isaacson, J. S., and Vitten, H. (2003). GABAB receptors inhibit dendrodendritic transmission in the rat olfactory bulb. J. Neurosci. 23, 2032–2039.

PubMed Abstract | Google Scholar

Jardemark, K., Nilsson, M., Muyderman, H., and Jacobson, I. (1997). Ca2+ ion permeability properties of (R,S) alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) receptors in isolated interneurons from the olfactory bulb of the rat. J. Neurophysiol. 77, 702–708.

PubMed Abstract | Google Scholar

Jeong, S.-W., Park, B.-G., Park, J.-Y., Lee, J.-W., and Lee, J.-H. (2003). Divalent metals differentially block cloned T-type calcium channels. Neuroreport 14, 1537–1540. doi: 10.1097/00001756-200308060-00028

PubMed Abstract | CrossRef Full Text | Google Scholar

Jo, S. M., Won, M. H., Cole, T. B., Jensen, M. S., Palmiter, R. D., and Danscher, G. (2000). Zinc-enriched (ZEN) terminals in mouse olfactory bulb. Brain Res. 865, 227–236. doi: 10.1016/s0006-8993(00)02227-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Johnston, J., and Delaney, K. R. (2010). Synaptic activation of T-type Ca2+ channels via mGluR activation in the primary dendrite of mitral cells. J. Neurophysiol. 103, 2557–2569. doi: 10.1152/jn.00796.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, M. V., and Westbrook, G. L. (1995). Desensitized states prolong GABAA channel responses to brief agonist pulses. Neuron 15, 181–191. doi: 10.1016/0896-6273(95)90075-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Kalappa, B. I., Anderson, C. T., Goldberg, J. M., Lippard, S. J., and Tzounopoulos, T. (2015). AMPA receptor inhibition by synaptically released zinc. Proc. Natl. Acad. Sci. U S A 112, 15749–15754. doi: 10.1073/pnas.1512296112

PubMed Abstract | CrossRef Full Text | Google Scholar

Kerchner, G. A., Canzoniero, L. M., Yu, S. P., Ling, C., and Choi, D. W. (2000). Zn2+ current is mediated by voltage-gated Ca2+ channels and enhanced by extracellular acidity in mouse cortical neurones. J. Physiol. 528, 39–52. doi: 10.1111/j.1469-7793.2000.00039.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Koh, J.-Y., and Choi, D. W. (1994). Zinc toxicity on cultured cortical neurons: involvement of N-methyl-D-aspartate receptors. Neuroscience 60, 1049–1057. doi: 10.1016/0306-4522(94)90282-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Kollo, M., Holderith, N., Antal, M., and Nusser, Z. (2008). Unique clustering of A-type potassium channels on different cell types of the main olfactory bulb. Eur. J. Neurosci. 27, 1686–1699. doi: 10.1111/j.1460-9568.2008.06141.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Kosaka, T., and Kosaka, K. (2016). Neuronal organization of the main olfactory bulb revisited. Anat. Sci. Int. 91, 115–127. doi: 10.1007/s12565-015-0309-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Kress, Y., Gaskin, F., Brosnan, C. F., and Levine, S. (1981). Effects of zinc on the cytoskeletal proteins in the central nervous system of the rat. Brain Res. 220, 139–149. doi: 10.1016/0006-8993(81)90217-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Kuo, C.-C., Chen, W.-Y., and Yang, Y.-C. (2004). Block of tetrodotoxin-resistant Na+ channel pore by multivalent cations: gating modification and Na+ flow dependence. J. Gen. Physiol. 124, 27–42. doi: 10.1085/jgp.200409054

PubMed Abstract | CrossRef Full Text | Google Scholar

Larson, A. A., and Kitto, K. F. (1997). Manipulations of zinc in the spinal cord, by intrathecal injection of zinc chloride, disodium-calcium-EDTA, or dipicolinic acid, alter nociceptive activity in mice. J. Pharmacol. Exp. Ther. 282, 1319–1325.

PubMed Abstract | Google Scholar

Larson, A. A., and Kitto, K. F. (1999). Chelation of zinc in the extracellular area of the spinal cord, using ethylenediaminetetraacetic acid disodium-calcium salt or dipicolinic acid, inhibits the antinociceptive effect of capsaicin in adult mice. J. Pharmacol. Exp. Ther. 288, 759–765.

PubMed Abstract | Google Scholar

Lassalle, J. M., Bataille, T., and Halley, H. (2000). Reversible inactivation of the hippocampal mossy fiber synapses in mice impairs spatial learning, but neither consolidation nor memory retrieval, in the Morris navigation task. Neurobiol. Learn. Mem. 73, 243–257. doi: 10.1006/nlme.1999.3931

PubMed Abstract | CrossRef Full Text | Google Scholar

Laube, B., Kuhse, J., and Betz, H. (2000). Kinetic and mutational analysis of Zn2+ modulation of recombinant human inhibitory glycine receptors. J. Physiol. 522, 215–230. doi: 10.1111/j.1469-7793.2000.t01-1-00215.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Laube, B., Kuhse, J., Rundström, N., Kirsch, J., Schmieden, V., and Betz, H. (1995). Modulation by zinc ions of native rat and recombinant human inhibitory glycine receptors. J. Physiol. 483, 613–619. doi: 10.1113/jphysiol.1995.sp020610

PubMed Abstract | CrossRef Full Text | Google Scholar

Laube, B., Maksay, G., Schemm, R., and Betz, H. (2002). Modulation of glycine receptor function: a novel approach for therapeutic intervention at inhibitory synapses?. Trends Pharmacol. Sci. 23, 519–527. doi: 10.1016/s0165-6147(02)02138-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Laurie, D. J., Seeburg, P. H., and Wisden, W. (1992). The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. II. Olfactory bulb and cerebellum. J. Neurosci. 12, 1063–1076.

PubMed Abstract | Google Scholar

Legendre, P., and Westbrook, G. L. (1990). The inhibition of single N-methyl-D-aspartate-activated channels by zinc ions on cultured rat neurones. J. Physiol. 429, 429–449. doi: 10.1113/jphysiol.1990.sp018266

PubMed Abstract | CrossRef Full Text | Google Scholar

Legendre, P., and Westbrook, G. L. (1991). Noncompetitive inhibition of gamma-aminobutyric acidA channels by Zn. Mol. Pharmacol. 39, 267–274.

PubMed Abstract | Google Scholar

Li, Y., Hough, C. J., Suh, S. W., Sarvey, J. M., and Frederickson, C. J. (2001). Rapid translocation of Zn2+ from presynaptic terminals into postsynaptic hippocampal neurons after physiological stimulation. J. Neurophysiol. 86, 2597–2604.

PubMed Abstract | Google Scholar

Lin, D. D., Cohen, A. S., and Coulter, D. A. (2001). Zinc-induced augmentation of excitatory synaptic currents and glutamate receptor responses in hippocampal CA3 neurons. J. Neurophysiol. 85, 1185–1196.

PubMed Abstract | Google Scholar

Liu, C. J., Grandes, P., Matute, C., Cuénod, M., and Streit, P. (1989). Glutamate-like immunoreactivity revealed in rat olfactory bulb, hippocampus and cerebellum by monoclonal antibody and sensitive staining method. Histochemistry 90, 427–445. doi: 10.1007/bf00494354

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Plachez, C., Shao, Z., Puche, A., and Shipley, M. T. (2013). Olfactory bulb short axon cell release of GABA and dopamine produces a temporally biphasic inhibition-excitation response in external tufted cells. J. Neurosci. 33, 2916–2926. doi: 10.1523/jneurosci.3607-12.2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Puche, A. C., and Shipley, M. T. (2016). The interglomerular circuit potently inhibits olfactory bulb output neurons by both direct and indirect pathways. J. Neurosci. 36, 9604–9617. doi: 10.1523/jneurosci.1763-16.2016

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., and Shipley, M. T. (2008a). Intrinsic conductances actively shape excitatory and inhibitory postsynaptic responses in olfactory bulb external tufted cells. J. Neurosci. 28, 10311–10322. doi: 10.1523/jneurosci.2608-08.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., and Shipley, M. T. (2008b). Multiple conductances cooperatively regulate spontaneous bursting in mouse olfactory bulb external tufted cells. J. Neurosci. 28, 1625–1639. doi: 10.1523/jneurosci.3906-07.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Lorincz, A., and Nusser, Z. (2008). Cell-type-dependent molecular composition of the axon initial segment. J. Neurosci. 28, 14329–14340. doi: 10.1523/jneurosci.4833-08.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Low, C. M., Zheng, F., Lyuboslavsky, P., and Traynelis, S. F. (2000). Molecular determinants of coordinated proton and zinc inhibition of N-methyl-D-aspartate NR1/NR2A receptors. Proc. Natl. Acad. Sci. U S A 97, 11062–11067. doi: 10.1073/pnas.180307497

PubMed Abstract | CrossRef Full Text | Google Scholar

Lynch, J. W. (2004). Molecular structure and function of the glycine receptor chloride channel. Physiol. Rev. 84, 1051–1095. doi: 10.1152/physrev.00042.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Lynch, J. W., Jacques, P., Pierce, K. D., and Schofield, P. R. (1998). Zinc potentiation of the glycine receptor chloride channel is mediated by allosteric pathways. J. Neurochem. 71, 2159–2168. doi: 10.1046/j.1471-4159.1998.71052159.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ma, J., and Lowe, G. (2007). Calcium permeable AMPA receptors and autoreceptors in external tufted cells of rat olfactory bulb. Neuroscience 144, 1094–1108. doi: 10.1016/j.neuroscience.2006.10.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Magistretti, J., Castelli, L., Taglietti, V., and Tanzi, F. (2003). Dual effect of Zn2+ on multiple types of voltage-dependent Ca2+ currents in rat palaeocortical neurons. Neuroscience 117, 249–264. doi: 10.1016/s0306-4522(02)00865-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Maher, B. J., and Westbrook, G. L. (2005). SK channel regulation of dendritic excitability and dendrodendritic inhibition in the olfactory bulb. J. Neurophysiol. 94, 3743–3750. doi: 10.1152/jn.00797.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Mandava, P., Howell, G. A., and Frederickson, C. J. (1993). Zinc-containing neuronal innervation of the septal nuclei. Brain Res. 608, 115–122. doi: 10.1016/0006-8993(93)90781-h

PubMed Abstract | CrossRef Full Text | Google Scholar

Martel, G., Hevi, C., Friebely, O., Baybutt, T., and Shumyatsky, G. P. (2010). Zinc transporter 3 is involved in learned fear and extinction, but not in innate fear. Learn. Mem. 17, 582–590. doi: 10.1101/lm.1962010

PubMed Abstract | CrossRef Full Text | Google Scholar

Martel, G., Hevi, C., Kane-Goldsmith, N., and Shumyatsky, G. P. (2011). Zinc transporter ZnT3 is involved in memory dependent on the hippocampus and perirhinal cortex. Behav. Brain Res. 223, 233–238. doi: 10.1016/j.bbr.2011.04.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Martin, L. J., Blackstone, C. D., Levey, A. I., Huganir, R. L., and Price, D. L. (1993). Cellular localizations of AMPA glutamate receptors within the basal forebrain magnocellular complex of rat and monkey. J. Neurosci. 13, 2249–2263.

PubMed Abstract | Google Scholar

Masters, B. A., Quaife, C. J., Erickson, J. C., Kelly, E. J., Froelick, G. J., Zambrowicz, B. P., et al. (1994). Metallothionein III is expressed in neurons that sequester zinc in synaptic vesicles. J. Neurosci. 14, 5844–5857.

PubMed Abstract | Google Scholar

Mathie, A., Sutton, G. L., Clarke, C. E., and Veale, E. L. (2006). Zinc and copper: pharmacological probes and endogenous modulators of neuronal excitability. Pharmacol. Ther. 111, 567–583. doi: 10.1016/j.pharmthera.2005.11.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Mayer, M. L., and Vyklicky, L. Jr. (1989). The action of zinc on synaptic transmission and neuronal excitability in cultures of mouse hippocampus. J. Physiol. 415, 351–365. doi: 10.1113/jphysiol.1989.sp017725

PubMed Abstract | CrossRef Full Text | Google Scholar

Mayer, M. L., Vyklicky, L. Jr., and Westbrook, G. L. (1989). Modulation of excitatory amino acid receptors by group IIB metal cations in cultured mouse hippocampal neurones. J. Physiol. 415, 329–350. doi: 10.1113/jphysiol.1989.sp017724

PubMed Abstract | CrossRef Full Text | Google Scholar

McAllister, B. B., and Dyck, R. H. (2017). Zinc transporter 3 (ZnT3) and vesicular zinc in central nervous system function. Neurosci. Biobehav. Rev. 80, 329–350. doi: 10.1016/j.neubiorev.2017.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

McCord, M. C., and Aizenman, E. (2013). Convergent Ca2+ and Zn2+ signaling regulates apoptotic Kv2.1 K+ currents. Proc. Natl. Acad. Sci. U S A 110, 13988–13993. doi: 10.1073/pnas.1306238110

PubMed Abstract | CrossRef Full Text | Google Scholar

McLaughlin, B., Pal, S., Tran, M. P., Parsons, A. A., Barone, F. C., Erhardt, J. A., et al. (2001). p38 activation is required upstream of potassium current enhancement and caspase cleavage in thiol oxidant-induced neuronal apoptosis. J. Neurosci. 21, 3303–3311.

PubMed Abstract | Google Scholar

Miller, P. S., Beato, M., Harvey, R. J., and Smart, T. G. (2005a). Molecular determinants of glycine receptor alphabeta subunit sensitivities to Zn2+-mediated inhibition. J. Physiol. 566, 657–670. doi: 10.1113/jphysiol.2005.088575

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, P. S., Da Silva, H. M., and Smart, T. G. (2005b). Molecular basis for zinc potentiation at strychnine-sensitive glycine receptors. J. Biol. Chem. 280, 37877–37884. doi: 10.1074/jbc.M508303200

PubMed Abstract | CrossRef Full Text | Google Scholar

Minami, A., Sakurada, N., Fuke, S., Kikuchi, K., Nagano, T., Oku, N., et al. (2006). Inhibition of presynaptic activity by zinc released from mossy fiber terminals during tetanic stimulation. J. Neurosci. Res. 83, 167–176. doi: 10.1002/jnr.20714

PubMed Abstract | CrossRef Full Text | Google Scholar

Molnár, E., Baude, A., Richmond, S. A., Patel, P. B., Somogyi, P., and McIlhinney, R. A. (1993). Biochemical and immunocytochemical characterization of antipeptide antibodies to a cloned GluR1 glutamate receptor subunit: cellular and subcellular distribution in the rat forebrain. Neuroscience 53, 307–326. doi: 10.1016/0306-4522(93)90198-o

PubMed Abstract | CrossRef Full Text | Google Scholar

Molnár, P., and Nadler, J. V. (2001). Synaptically-released zinc inhibits N-methyl-D-aspartate receptor activation at recurrent mossy fiber synapses. Brain Res. 910, 205–207. doi: 10.1016/s0006-8993(01)02720-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Montague, A. A., and Greer, C. A. (1999). Differential distribution of ionotropic glutamate receptor subunits in the rat olfactory bulb. J. Comp. Neurol. 405, 233–246. doi: 10.1002/(sici)1096-9861(19990308)405:2<233::aid-cne7>3.0.co;2-a

PubMed Abstract | CrossRef Full Text | Google Scholar

Monyer, H., Burnashev, N., Laurie, D. J., Sakmann, B., and Seeburg, P. H. (1994). Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron 12, 529–540. doi: 10.1016/0896-6273(94)90210-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Mook Jo, S., Kuk Kim, Y., Wang, Z., and Danscher, G. (2002). Retrograde tracing of zinc-enriched (ZEN) neuronal somata projecting to the olfactory bulb. Brain Res. 956, 230–235. doi: 10.1016/s0006-8993(02)03544-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Mori, K., Kishi, K., and Ojima, H. (1983). Distribution of dendrites of mitral, displaced mitral, tufted, and granule cells in the rabbit olfactory bulb. J. Comp. Neurol. 219, 339–355. doi: 10.1002/cne.902190308

PubMed Abstract | CrossRef Full Text | Google Scholar

Nagayama, S., Homma, R., and Imamura, F. (2014). Neuronal organization of olfactory bulb circuits. Front. Neural Circuits 8:98. doi: 10.3389/fncir.2014.00098

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakashima, A. S., and Dyck, R. H. (2009). Zinc and cortical plasticity. Brain Res. Rev. 59, 347–373. doi: 10.1016/j.brainresrev.2008.10.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Nevin, S. T., Cromer, B. A., Haddrill, J. L., Morton, C. J., Parker, M. W., and Lynch, J. W. (2003). Insights into the structural basis for zinc inhibition of the glycine receptor. J. Biol. Chem. 278, 28985–28992. doi: 10.1074/jbc.M300097200

PubMed Abstract | CrossRef Full Text | Google Scholar

Nusser, Z. (2009). Variability in the subcellular distribution of ion channels increases neuronal diversity. Trends Neurosci. 32, 267–274. doi: 10.1016/j.tins.2009.01.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Ono, S., and Cherian, M. G. (1999). Regional distribution of metallothionein, zinc, and copper in the brain of different strains of rats. Biol. Trace Elem. Res. 69, 151–159. doi: 10.1007/bf02783866

PubMed Abstract | CrossRef Full Text | Google Scholar

Pal, S., Hartnett, K. A., Nerbonne, J. M., Levitan, E. S., and Aizenman, E. (2003). Mediation of neuronal apoptosis by Kv2.1-encoded potassium channels. J. Neurosci. 23, 4798–4802.

PubMed Abstract | Google Scholar

Pal, S. K., Takimoto, K., Aizenman, E., and Levitan, E. S. (2006). Apoptotic surface delivery of K+ channels. Cell Death Differ. 13, 661–667. doi: 10.1038/sj.cdd.4401792

PubMed Abstract | CrossRef Full Text | Google Scholar

Palmiter, R. D., Cole, T. B., Quaife, C. J., and Findley, S. D. (1996). ZnT-3, a putative transporter of zinc into synaptic vesicles. Proc. Natl. Acad. Sci. U S A 93, 14934–14939. doi: 10.1073/pnas.93.25.14934

PubMed Abstract | CrossRef Full Text | Google Scholar

Pan, E., Zhang, X. A., Huang, Z., Krezel, A., Zhao, M., Tinberg, C. E., et al. (2011). Vesicular zinc promotes presynaptic and inhibits postsynaptic long-term potentiation of mossy fiber-CA3 synapse. Neuron 71, 1116–1126. doi: 10.1016/j.neuron.2011.07.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Panzanelli, P., Perazzini, A. Z., Fritschy, J. M., and Sassoè-Pognetto, M. (2005). Heterogeneity of gamma-aminobutyric acid type A receptors in mitral and tufted cells of the rat main olfactory bulb. J. Comp. Neurol. 484, 121–131. doi: 10.1002/cne.20440

PubMed Abstract | CrossRef Full Text | Google Scholar

Paoletti, P., Ascher, P., and Neyton, J. (1997). High-affinity zinc inhibition of NMDA NR1-NR2A receptors. J. Neurosci. 17, 5711–5725.

PubMed Abstract | Google Scholar

Paoletti, P., Bellone, C., and Zhou, Q. (2013). NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci. 14, 383–400. doi: 10.1038/nrn3504

PubMed Abstract | CrossRef Full Text | Google Scholar

Paoletti, P., Vergnano, A. M., Barbour, B., and Casado, M. (2009). Zinc at glutamatergic synapses. Neuroscience 158, 126–136. doi: 10.1016/j.neuroscience.2008.01.061

PubMed Abstract | CrossRef Full Text | Google Scholar

Papke, D., Gonzalez-Gutierrez, G., and Grosman, C. (2011). Desensitization of neurotransmitter-gated ion channels during high-frequency stimulation: a comparative study of Cys-loop, AMPA and purinergic receptors. J. Physiol. 589, 1571–1585. doi: 10.1113/jphysiol.2010.203315

PubMed Abstract | CrossRef Full Text | Google Scholar

Pérez-Clausell, J. (1996). Distribution of terminal fields stained for zinc in the neocortex of the rat. J. Chem. Neuroanat. 11, 99–111. doi: 10.1016/0891-0618(96)00131-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Pérez-Clausell, J., and Danscher, G. (1985). Intravesicular localization of zinc in rat telencephalic boutons. A histochemical study. Brain Res. 337, 91–98. doi: 10.1016/0006-8993(85)91612-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Perez-Rosello, T., Anderson, C. T., Ling, C., Lippard, S. J., and Tzounopoulos, T. (2015). Tonic zinc inhibits spontaneous firing in dorsal cochlear nucleus principal neurons by enhancing glycinergic neurotransmission. Neurobiol. Dis. 81, 14–19. doi: 10.1016/j.nbd.2015.03.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Perez-Rosello, T., Anderson, C. T., Schopfer, F. J., Zhao, Y., Gilad, D., Salvatore, S. R., et al. (2013). Synaptic Zn2+ inhibits neurotransmitter release by promoting endocannabinoid synthesis. J. Neurosci. 33, 9259–9272. doi: 10.1523/JNEUROSCI.0237-13.2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Peters, S., Koh, J., and Choi, D. W. (1987). Zinc selectively blocks the action of N-methyl-D-aspartate on cortical neurons. Science 236, 589–593. doi: 10.1126/science.2883728

PubMed Abstract | CrossRef Full Text | Google Scholar

Petralia, R. S., and Wenthold, R. J. (1992). Light and electron immunocytochemical localization of AMPA-selective glutamate receptors in the rat brain. J. Comp. Neurol. 318, 329–354. doi: 10.1002/cne.903180309

PubMed Abstract | CrossRef Full Text | Google Scholar

Petralia, R. S., Yokotani, N., and Wenthold, R. J. (1994). Light and electron microscope distribution of the NMDA receptor subunit NMDAR1 in the rat nervous system using a selective anti-peptide antibody. J. Neurosci. 14, 667–696.

PubMed Abstract | Google Scholar

Pignatelli, A., Kobayashi, K., Okano, H., and Belluzzi, O. (2005). Functional properties of dopaminergic neurones in the mouse olfactory bulb. J. Physiol. 564, 501–514. doi: 10.1113/jphysiol.2005.084632

PubMed Abstract | CrossRef Full Text | Google Scholar

Pimentel, D. O., and Margrie, T. W. (2008). Glutamatergic transmission and plasticity between olfactory bulb mitral cells. J. Physiol. 586, 2107–2119. doi: 10.1113/jphysiol.2007.149575

PubMed Abstract | CrossRef Full Text | Google Scholar

Pinching, A. J., and Powell, T. P. (1971). The neuron types of the glomerular layer of the olfactory bulb. J. Cell Sci. 9, 305–345.

PubMed Abstract | Google Scholar

Pourcho, R. G., Goebel, D. J., Jojich, L., and Hazlett, J. C. (1992). Immunocytochemical evidence for the involvement of glycine in sensory centers of the rat brain. Neuroscience 46, 643–656. doi: 10.1016/0306-4522(92)90151-q

PubMed Abstract | CrossRef Full Text | Google Scholar

Price, J. L., and Powell, T. P. (1970a). The morphology of the granule cells of the olfactory bulb. J. Cell Sci. 7, 91–123.

PubMed Abstract | Google Scholar

Price, J. L., and Powell, T. P. (1970b). The synaptology of the granule cells of the olfactory bulb. J. Cell Sci. 7, 125–155.

PubMed Abstract | Google Scholar

Puopolo, M., and Belluzzi, O. (1998). Functional heterogeneity of periglomerular cells in the rat olfactory bulb. Eur. J. Neurosci. 10, 1073–1083. doi: 10.1046/j.1460-9568.1998.00115.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Qian, J., and Noebels, J. L. (2005). Visualization of transmitter release with zinc fluorescence detection at the mouse hippocampal mossy fibre synapse. J. Physiol. 566, 747–758. doi: 10.1113/jphysiol.2005.089276

PubMed Abstract | CrossRef Full Text | Google Scholar

Qian, J., and Noebels, J. L. (2006). Exocytosis of vesicular zinc reveals persistent depression of neurotransmitter release during metabotropic glutamate receptor long-term depression at the hippocampal CA3-CA1 synapse. J. Neurosci. 26, 6089–6095. doi: 10.1523/JNEUROSCI.0475-06.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Quinta-Ferreira, M. E., and Matias, C. M. (2004). Hippocampal mossy fiber calcium transients are maintained during long-term potentiation and are inhibited by endogenous zinc. Brain Res. 1004, 52–60. doi: 10.1016/j.brainres.2004.01.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Rachline, J., Perin-Dureau, F., Le Goff, A., Neyton, J., and Paoletti, P. (2005). The micromolar zinc-binding domain on the NMDA receptor subunit NR2B. J. Neurosci. 25, 308–317. doi: 10.1523/jneurosci.3967-04.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Rademakers, R., Neumann, M., and Mackenzie, I. R. (2012). Advances in understanding the molecular basis of frontotemporal dementia. Nat. Rev. Neurol. 8, 423–434. doi: 10.1038/nrneurol.2012.117

PubMed Abstract | CrossRef Full Text | Google Scholar

Rall, W., Shepherd, G. M., Reese, T. S., and Brightman, M. W. (1966). Dendrodendritic synaptic pathway for inhibition in the olfactory bulb. Exp. Neurol. 14, 44–56. doi: 10.1016/0014-4886(66)90023-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Rassendren, F. A., Lory, P., Pin, J. P., and Nargeot, J. (1990). Zinc has opposite effects on NMDA and non-NMDA receptors expressed in Xenopus oocytes. Neuron 4, 733–740. doi: 10.1016/0896-6273(90)90199-p

PubMed Abstract | CrossRef Full Text | Google Scholar

Redman, P. T., Hartnett, K. A., Aras, M. A., Levitan, E. S., and Aizenman, E. (2009). Regulation of apoptotic potassium currents by coordinated zinc-dependent signalling. J. Physiol. 587, 4393–4404. doi: 10.1113/jphysiol.2009.176321

PubMed Abstract | CrossRef Full Text | Google Scholar

Redman, P. T., He, K., Hartnett, K. A., Jefferson, B. S., Hu, L., Rosenberg, P. A., et al. (2007). Apoptotic surge of potassium currents is mediated by p38 phosphorylation of Kv2.1. Proc. Natl. Acad. Sci. U S A 104, 3568–3573. doi: 10.1073/pnas.0610159104

PubMed Abstract | CrossRef Full Text | Google Scholar

Rigo, J. M., and Legendre, P. (2006). Frequency-dependent modulation of glycine receptor activation recorded from the zebrafish larvae hindbrain. Neuroscience 140, 389–402. doi: 10.1016/j.neuroscience.2006.01.057

PubMed Abstract | CrossRef Full Text | Google Scholar

Ruiz, A., Walker, M. C., Fabian-Fine, R., and Kullmann, D. M. (2004). Endogenous zinc inhibits GABAA receptors in a hippocampal pathway. J. Neurophysiol. 91, 1091–1096. doi: 10.1152/jn.00755.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Schofield, P. R., Darlison, M. G., Fujita, N., Burt, D. R., Stephenson, F. A., Rodriguez, H., et al. (1987). Sequence and functional expression of the GABA A receptor shows a ligand-gated receptor super-family. Nature 328, 221–227. doi: 10.1038/328221a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoppa, N. E. (2006). Synchronization of olfactory bulb mitral cells by precisely timed inhibitory inputs. Neuron 49, 271–283. doi: 10.1016/j.neuron.2005.11.038

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoppa, N. E., Kinzie, J. M., Sahara, Y., Segerson, T. P., and Westbrook, G. L. (1998). Dendrodendritic inhibition in the olfactory bulb is driven by NMDA receptors. J. Neurosci. 18, 6790–6802.

PubMed Abstract | Google Scholar

Schoppa, N. E., and Westbrook, G. L. (1999). Regulation of synaptic timing in the olfactory bulb by an A-type potassium current. Nat. Neurosci. 2, 1106–1113. doi: 10.1038/16033

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoppa, N. E., and Westbrook, G. L. (2001). Glomerulus-specific synchronization of mitral cells in the olfactory bulb. Neuron 31, 639–651. doi: 10.1016/s0896-6273(01)00389-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoppa, N. E., and Westbrook, G. L. (2002). AMPA autoreceptors drive correlated spiking in olfactory bulb glomeruli. Nat. Neurosci. 5, 1194–1202. doi: 10.1038/nn953

PubMed Abstract | CrossRef Full Text | Google Scholar

Segal, D., Ohana, E., Besser, L., Hershfinkel, M., Moran, A., and Sekler, I. (2004). A role for ZnT-1 in regulating cellular cation influx. Biochem. Biophys. Res. Commun. 323, 1145–1150. doi: 10.1016/j.bbrc.2004.08.211

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekler, I., Moran, A., Hershfinkel, M., Dori, A., Margulis, A., Birenzweig, N., et al. (2002). Distribution of the zinc transporter ZnT-1 in comparison with chelatable zinc in the mouse brain. J. Comp. Neurol. 447, 201–209. doi: 10.1002/cne.10224

PubMed Abstract | CrossRef Full Text | Google Scholar

Sensi, S. L., Canzoniero, L. M., Yu, S. P., Ying, H. S., Koh, J. Y., Kerchner, G. A., et al. (1997). Measurement of intracellular free zinc in living cortical neurons: routes of entry. J. Neurosci. 17, 9554–9564.

PubMed Abstract | Google Scholar

Sensi, S. L., Yin, H. Z., Carriedo, S. G., Rao, S. S., and Weiss, J. H. (1999). Preferential Zn2+ influx through Ca2+-permeable AMPA/kainate channels triggers prolonged mitochondrial superoxide production. Proc. Natl. Acad. Sci. U S A 96, 2414–2419. doi: 10.1073/pnas.96.5.2414

PubMed Abstract | CrossRef Full Text | Google Scholar

Serafini, R., Valeyev, A. Y., Barker, J. L., and Poulter, M. O. (1995). Depolarizing GABA-activated Cl- channels in embryonic rat spinal and olfactory bulb cells. J. Physiol. 488, 371–386. doi: 10.1113/jphysiol.1995.sp020973

PubMed Abstract | CrossRef Full Text | Google Scholar

Shen, Y., and Yang, X. L. (1999). Zinc modulation of AMPA receptors may be relevant to splice variants in carp retina. Neurosci. Lett. 259, 177–180. doi: 10.1016/s0304-3940(98)00938-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Shepherd, G. M. (1972). Synaptic organization of the mammalian olfactory bulb. Physiol. Rev. 52, 864–917.

PubMed Abstract | Google Scholar

Shepherd, G. M., Chen, W. R., and Greer, C. A. (2004). “Olfactory bulb,” in The Synaptic Organization of the Brain, 5th Edn. ed. G. M. Shepherd (New York, NY: Oxford University Press), 165–216.

Google Scholar

Shusterman, E., Beharier, O., Shiri, L., Zarivach, R., Etzion, Y., Campbell, C. R., et al. (2014). ZnT-1 extrudes zinc from mammalian cells functioning as a Zn2+/H+ exchanger. Metallomics 6, 1656–1663. doi: 10.1039/c4mt00108g

PubMed Abstract | CrossRef Full Text | Google Scholar

Sindreu, C., Palmiter, R. D., and Storm, D. R. (2011). Zinc transporter ZnT-3 regulates presynaptic Erk1/2 signaling and hippocampus-dependent memory. Proc. Natl. Acad. Sci. U S A 108, 3366–3370. doi: 10.1073/pnas.1019166108

PubMed Abstract | CrossRef Full Text | Google Scholar

Slomianka, L. (1992). Neurons of origin of zinc-containing pathways and the distribution of zinc-containing boutons in the hippocampal region of the rat. Neuroscience 48, 325–352. doi: 10.1016/0306-4522(92)90494-m

PubMed Abstract | CrossRef Full Text | Google Scholar

Slomianka, L., Danscher, G., and Frederickson, C. J. (1990). Labeling of the neurons of origin of zinc-containing pathways by intraperitoneal injections of sodium selenite. Neuroscience 38, 843–854. doi: 10.1016/0306-4522(90)90076-g

PubMed Abstract | CrossRef Full Text | Google Scholar

Smart, T. G., Xie, X., and Krishek, B. J. (1994). Modulation of inhibitory and excitatory amino acid receptor ion channels by zinc. Prog. Neurobiol. 42, 393–441. doi: 10.1016/0301-0082(94)90082-5

PubMed Abstract | CrossRef Full Text

Takahashi, K., and Akaike, N. (1991). Calcium antagonist effects on low-threshold (T-type) calcium current in rat isolated hippocampal CA1 pyramidal neurons. J. Pharmacol. Exp. Ther. 256, 169–175.

PubMed Abstract | Google Scholar

Takeda, A., Fuke, S., Ando, M., and Oku, N. (2009). Positive modulation of long-term potentiation at hippocampal CA1 synapses by low micromolar concentrations of zinc. Neuroscience 158, 585–591. doi: 10.1016/j.neuroscience.2008.10.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Takeda, A., Fuke, S., Tsutsumi, W., and Oku, N. (2007). Negative modulation of presynaptic activity by zinc released from Schaffer collaterals. J. Neurosci. Res. 85, 3666–3672. doi: 10.1002/jnr.21449

PubMed Abstract | CrossRef Full Text | Google Scholar

Takeda, A., Tamano, H., Imano, S., and Oku, N. (2010). Increases in extracellular zinc in the amygdala in acquisition and recall of fear experience and their roles in response to fear. Neuroscience 168, 715–722. doi: 10.1016/j.neuroscience.2010.04.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Talley, E. M., Cribbs, L. L., Lee, J. H., Daud, A., Perez-Reyes, E., and Bayliss, D. A. (1999). Differential distribution of three members of a gene family encoding low voltage-activated (T-type) calcium channels. J. Neurosci. 19, 1895–1911.

PubMed Abstract | Google Scholar

Tamano, H., and Takeda, A. (2011). Dynamic action of neurometals at the synapse. Metallomics 3, 656–661. doi: 10.1039/c1mt00008j

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, H. B., Miyano, K., and Nakata, Y. (2009). Modulation of the substance P release from cultured rat primary afferent neurons by zinc ions. J. Pharmacol. Sci. 110, 397–400. doi: 10.1254/jphs.09033sc

PubMed Abstract | CrossRef Full Text | Google Scholar

Trombley, P. Q., Blakemore, L. J., and Hill, B. J. (2011). Zinc modulation of glycine receptors. Neuroscience 186, 32–38. doi: 10.1016/j.neuroscience.2011.04.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Trombley, P. Q., Hill, B. J., and Horning, M. S. (1999). Interactions between GABA and glycine at inhibitory amino acid receptors on rat olfactory bulb neurons. J. Neurophysiol. 82, 3417–3422.

PubMed Abstract | Google Scholar

Trombley, P. Q., Horning, M. S., and Blakemore, L. J. (1998). Carnosine modulates zinc and copper effects on amino acid receptors and synaptic transmission. Neuroreport 9, 3503–3507. doi: 10.1097/00001756-199810260-00031

PubMed Abstract | CrossRef Full Text | Google Scholar

Trombley, P. Q., and Shepherd, G. M. (1993). Synaptic transmission and modulation in the olfactory bulb. Curr. Opin. Neurobiol. 3, 540–547. doi: 10.1016/0959-4388(93)90053-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Trombley, P. Q., and Shepherd, G. M. (1994). Glycine exerts potent inhibitory actions on mammalian olfactory bulb neurons. J. Neurophysiol. 71, 761–767.

PubMed Abstract | Google Scholar

Trombley, P. Q., and Shepherd, G. M. (1996). Differential modulation by zinc and copper of amino acid receptors from rat olfactory bulb neurons. J. Neurophysiol. 76, 2536–2546.

PubMed Abstract | Google Scholar

Trombley, P. Q., and Westbrook, G. L. (1990). Excitatory synaptic transmission in cultures of rat olfactory bulb. J. Neurophysiol. 64, 598–606.

PubMed Abstract | Google Scholar

Ueno, S., Tsukamoto, M., Hirano, T., Kikuchi, K., Yamada, M. K., Nishiyama, N., et al. (2002). Mossy fiber Zn2+ spillover modulates heterosynaptic N-methyl-D-aspartate receptor activity in hippocampal CA3 circuits. J. Cell Biol. 158, 215–220. doi: 10.1083/jcb.200204066

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaaga, C. E., Yorgason, J. T., Williams, J. T., and Westbrook, G. L. (2017). Presynaptic gain control by endogenous cotransmission of dopamine and GABA in the olfactory bulb. J. Neurophysiol. 117, 1163–1170. doi: 10.1152/jn.00694.2016

PubMed Abstract | CrossRef Full Text | Google Scholar

van den Pol, A. N. (1995). Presynaptic metabotropic glutamate receptors in adult and developing neurons: autoexcitation in the olfactory bulb. J. Comp. Neurol. 359, 253–271. doi: 10.1002/cne.903590206

PubMed Abstract | CrossRef Full Text | Google Scholar

van den Pol, A. N., and Gorcs, T. (1988). Glycine and glycine receptor immunoreactivity in brain and spinal cord. J. Neurosci. 8, 472–492.

PubMed Abstract | Google Scholar

Veh, R. W., Lichtinghagen, R., Sewing, S., Wunder, F., Grumbach, I. M., and Pongs, O. (1995). Immunohistochemical localization of five members of the Kv1 channel subunits: contrasting subcellular locations and neuron-specific co-localizations in rat brain. Eur. J. Neurosci. 7, 2189–2205. doi: 10.1111/j.1460-9568.1995.tb00641.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Verdoorn, T. A., Burnashev, N., Monyer, H., Seeburg, P. H., and Sakmann, B. (1991). Structural determinants of ion flow through recombinant glutamate receptor channels. Science 252, 1715–1718. doi: 10.1126/science.1710829

PubMed Abstract | CrossRef Full Text | Google Scholar

Vergnano, A. M., Rebola, N., Savtchenko, L. P., Pinheiro, P. S., Casado, M., Kieffer, B. L., et al. (2014). Zinc dynamics and action at excitatory synapses. Neuron 82, 1101–1114. doi: 10.1016/j.neuron.2014.04.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Vogt, K., Mellor, J., Tong, G., and Nicoll, R. (2000). The actions of synaptically released zinc at hippocampal mossy fiber synapses. Neuron 26, 187–196. doi: 10.1016/s0896-6273(00)81149-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Danscher, G., Kim, Y. K., Dahlstrom, A., and Mook Jo, S. (2002). Inhibitory zinc-enriched terminals in the mouse cerebellum: double-immunohistochemistry for zinc transporter 3 and glutamate decarboxylase. Neurosci. Lett. 321, 37–40. doi: 10.1016/s0304-3940(01)02560-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Li, J. Y., Dahlström, A., and Danscher, G. (2001). Zinc-enriched GABAergic terminals in mouse spinal cord. Brain Res. 921, 165–172. doi: 10.1016/s0006-8993(01)03114-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Watanabe, M., Inoue, Y., Sakimura, K., and Mishina, M. (1993). Distinct distributions of five N-methyl-D-aspartate receptor channel subunit mRNAs in the forebrain. J. Comp. Neurol. 338, 377–390. doi: 10.1002/cne.903380305

PubMed Abstract | CrossRef Full Text | Google Scholar

Weiss, J. H., and Sensi, S. L. (2000). Ca2+-Zn2+ permeable AMPA or kainate receptors: possible key factors in selective neurodegeneration. Trends Neurosci. 23, 365–371. doi: 10.1016/s0166-2236(00)01610-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Weltzien, F., Puller, C., O’Sullivan, G. A., Paarmann, I., and Betz, H. (2012). Distribution of the glycine receptor β-subunit in the mouse CNS as revealed by a novel monoclonal antibody. J. Comp. Neurol. 520, 3962–3981. doi: 10.1002/cne.23139

PubMed Abstract | CrossRef Full Text | Google Scholar

Westbrook, G. L., and Mayer, M. L. (1987). Micromolar concentrations of Zn2+ antagonize NMDA and GABA responses of hippocampal neurons. Nature 328, 640–643. doi: 10.1038/328640a0

PubMed Abstract | CrossRef Full Text | Google Scholar

White, J. A., Alonso, A., and Kay, A. R. (1993). A heart-like Na+ current in the medial entorhinal cortex. Neuron 11, 1037–1047. doi: 10.1016/0896-6273(93)90217-f

PubMed Abstract | CrossRef Full Text | Google Scholar

Woolf, T. B., Shepherd, G. M., and Greer, C. A. (1991). Serial reconstructions of granule cell spines in the mammalian olfactory bulb. Synapse 7, 181–192. doi: 10.1002/syn.890070303

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, X., Gerber, U., Gähwiler, B. H., and Smart, T. G. (1993). Interaction of zinc with ionotropic and metabotropic glutamate receptors in rat hippocampal slices. Neurosci. Lett. 159, 46–50. doi: 10.1016/0304-3940(93)90795-m

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, X. M., and Smart, T. G. (1991). A physiological role for endogenous zinc in rat hippocampal synaptic neurotransmission. Nature 349, 521–524. doi: 10.1038/349521a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Yin, H. Z., and Weiss, J. H. (1995). Zn2+ permeates Ca2+ permeable AMPA/kainate channels and triggers selective neural injury. Neuroreport 6, 2553–2556. doi: 10.1097/00001756-199512150-00025

PubMed Abstract | CrossRef Full Text | Google Scholar

Yoo, M. H., Kim, T. Y., Yoon, Y. H., and Koh, J. Y. (2016). Autism phenotypes in ZnT3 null mice: involvement of zinc dyshomeostasis, MMP-9 activation and BDNF upregulation. Sci. Rep. 6:28548. doi: 10.1038/srep28548

PubMed Abstract | CrossRef Full Text | Google Scholar

Youngentob, S. L., Mozell, M. M., Sheehe, P. R., and Hornung, D. E. (1987). A quantitative analysis of sniffing strategies in rats performing odor detection tasks. Physiol. Behav. 41, 59–69. doi: 10.1016/0031-9384(87)90131-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X. A., Lovejoy, K. S., Jasanoff, A., and Lippard, S. J. (2007). Water-soluble porphyrins as a dual-function molecular imaging platform for MRI and fluorescence zinc sensing. Proc. Natl. Acad. Sci. U S A 104, 10780–10785. doi: 10.1073/pnas.0702393104

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: zinc, olfactory bulb, central nervous system, modulation, amino acid receptors, voltage-gated ion channels

Citation: Blakemore LJ and Trombley PQ (2017) Zinc as a Neuromodulator in the Central Nervous System with a Focus on the Olfactory Bulb. Front. Cell. Neurosci. 11:297. doi: 10.3389/fncel.2017.00297

Received: 19 July 2017; Accepted: 06 September 2017;
Published: 21 September 2017.

Edited by:

Qi Yuan, Memorial University of Newfoundland, Canada

Reviewed by:

Alberto Granzotto, CeSI-MeT - Centro Scienze dell’Invecchiamento e Medicina Traslazionale, Italy
Richard H. Dyck, University of Calgary, Canada

Copyright © 2017 Blakemore and Trombley. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Paul Q. Trombley, trombley@neuro.fsu.edu

Download