Next Article in Journal
Regulation of Chemokine–Receptor Interactions and Functions
Next Article in Special Issue
Zinc Signal in Brain Diseases
Previous Article in Journal
Targeting Immune Cell Checkpoints during Sepsis
Previous Article in Special Issue
Impact of Labile Zinc on Heart Function: From Physiology to Pathophysiology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Impact of Synaptic Zn2+ Dynamics on Cognition and Its Decline

Department of Neurophysiology, School of Pharmaceutical Sciences, University of Shizuoka, 52-1 Yada, Suruga-ku, Shizuoka 422-8526, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(11), 2411; https://doi.org/10.3390/ijms18112411
Submission received: 22 September 2017 / Revised: 31 October 2017 / Accepted: 9 November 2017 / Published: 14 November 2017
(This article belongs to the Special Issue Zinc Signaling in Physiology and Pathogenesis)

Abstract

:
The basal levels of extracellular Zn2+ are in the range of low nanomolar concentrations and less attention has been paid to Zn2+, compared to Ca2+, for synaptic activity. However, extracellular Zn2+ is necessary for synaptic activity. The basal levels of extracellular zinc are age-dependently increased in the rat hippocampus, implying that the basal levels of extracellular Zn2+ are also increased age-dependently and that extracellular Zn2+ dynamics are linked with age-related cognitive function and dysfunction. In the hippocampus, the influx of extracellular Zn2+ into postsynaptic neurons, which is often linked with Zn2+ release from neuron terminals, is critical for cognitive activity via long-term potentiation (LTP). In contrast, the excess influx of extracellular Zn2+ into postsynaptic neurons induces cognitive decline. Interestingly, the excess influx of extracellular Zn2+ more readily occurs in aged dentate granule cells and intracellular Zn2+-buffering, which is assessed with ZnAF-2DA, is weakened in the aged dentate granule cells. Characteristics (easiness) of extracellular Zn2+ influx seem to be linked with the weakened intracellular Zn2+-buffering in the aged dentate gyrus. This paper deals with the impact of synaptic Zn2+ signaling on cognition and its decline in comparison with synaptic Ca2+ signaling.

Graphical Abstract

1. Introduction

Cognitive activity has been closely linked to strengthening and weakening synaptic connections between neurons that is synaptic plasticity such as long-term potentiation (LTP) and long-term depression (LTD). The hippocampal formation, which spans the posterior-to-anterior extent of the base of the temporal lobes, plays a key role in learning, memory, and recognition of novelty [1]. In its transverse axis, the hippocampal formation consists of the entorhinal cortex, the dentate gyrus, the CA3 and the CA1 subfields, and the subiculum. The entorhinal cortex functions as the gateway into the hippocampal formation [2]. The entorhinal cortex layer II projects to dentate granule cells via the perforant pathway, and dentate granule cells project to CA3 pyramidal cells via the mossy fibers. CA3 pyramidal cells interconnect with other CA3 neurons and project to CA1 pyramidal cells via the Shaffer collaterals. Finally, CA1 pyramidal cells connect to the subiculum. The entorhinal cortex layer II also projects to CA3 pyramidal cells, and the entorhinal cortex layer III also projects to CA1 pyramidal cells and the subiculum [3].
In the process of changes in synaptic structure for memory formation, glutamatergic neurons play a key role in the main neural circuit of the hippocampal formation. Research on synaptic plasticity opens a window for the molecular mechanisms of memory. Changes in both presynaptic and postsynaptic strength have been implicated in the mechanisms of LTP and LTD, and attention has been paid to changes in postsynaptic glutamate receptor density [4]. Intracellular Ca2+ signaling via the network of signaling molecules controls the glutamate receptor density and induces synaptic signals, as is observed in learning situations [5], followed by cognitive performance and memory.
For N-methyl-d-aspartate (NMDA)-receptor-dependent plasticity, the influx of extracellular Ca2+ into postsynaptic neurons through NMDA receptors plays a key role [6]. However, glutamate receptor activation by excess of extracellular glutamate, which is known as glutamate excitotoxicity [7,8], leads to a final common pathway for neuronal death and is linked with pathophysiological processes of neurological disorders [9,10]. CA1 pyramidal cells are the most vulnerable to neurodegeneration in the hippocampus after stroke/ischemia [11,12,13]. A well-known fact is that extracellular Ca2+ influx into postsynaptic neurons, in addition to Ca2+ release from the calcium stores results in neuronal death.
On the other hand, extracellular glutamate signaling also induces cellular transients in Zn2+ concentration, i.e., intracellular Zn2+ signaling, which is required for synaptic plasticity [14,15] and may have crosstalk to intracellular Ca2+ signaling via calcium channels [16,17]. On the basis of the subsequent evidence that glutamate-induced neuronal death is due to extracellular Zn2+ influx into postsynaptic neurons, which is dynamically linked with Zn2+ release from zincergic neurons, a subclass of glutamatergic neurons that concentrate zinc in the presynaptic vesicles [18,19,20,21,22], this paper deals with the impact of synaptic Zn2+ signaling on cognitive function and dysfunction in comparison with synaptic Ca2+ signaling. While the hydrated Ca2+ ion is the major species in intracellular Ca2+ signaling, this is not the case in intracellular Zn2+ signaling because the Zn2+ ion has much higher affinities for donors of ligands [23]. The Zn2+ ion is different from the Ca2+ and Mg2+ ions because it forms much stronger complexes with water and various anions and ligands. These characteristics are important for its synaptic functions.

2. Physiology of Brain Zn2+

Divalent cations such as Ca2+ and Mg2+ are involved in synaptic neurotransmission [24]. Among divalent cations, Ca2+ concentration is the highest in brain parenchyma cells and is approximately 1.2 mM in the cerebrospinal fluid (CSF) and brain extracellular fluid in the adult rats (Figure 1) [25]. Approximately 2 mM Ca2+ is added to artificial cerebrospinal fluid (ACSF) based on essentiality of intracellular Ca2+ signaling in neurons and glial cells [26,27]. However, excess influx of extracellular Ca2+ into neurons is linked with the pathophysiological process of neurodegeneration [28,29,30].
Zinc concentration in the CSF is in the range of 150–380 nM [31,32,33]. It is estimated that the basal (static) concentration of extracellular Zn2+ is approximately 10 nM in the brain of the adult rats (Figure 1) [34]. A small part of extracellular zinc is free ion (Zn2+) in the brain under the basal condition. To research synaptic function, much less attention has been paid to the essentiality of Zn2+ in brain extracellular fluid. ACSF, i.e., brain extracellular medium, without Zn2+ has been used for in vitro and in vivo experiments. It is likely that not only neuronal excitation but also LTP is modified in brain slices immersed in ACSF without Zn2+, in which original neurophysiology might be modified [35,36]. Clarifying the action of extracellular Zn2+ in the range of physiological concentrations is important to precisely understand synaptic function. Furthermore, such clarification is also important to understand the bidirectional action of Zn2+ under physiological and pathological conditions. It is recognized that low nanomolar concentrations of Zn2+ are more physiological than micromolar concentrations of Zn2+, which are widely used and often neurotoxic.
Spontaneous presynaptic activity assessed with FM4-64, an indicator of presynaptic activity (exocytosis), in the stratum lucidum where mossy fibers are contained is significantly suppressed in brain slices from young rats immersed in ACSF containing 10 nM Zn2+, but not in ACSF containing 10 nM Cu2+ or 10 nM Fe3+, indicating that hippocampal presynaptic activity is enhanced in brain slices prepared with ACSF without Zn2+ [36]. Suh et al. [37] report that acute brain slice preparations are poorly suitable to research the role of endogenous Zn2+ released from zincergic neurons. Vesicular Zn2+ levels are decreased in the process of slice preparation, and in vitro Zn2+ release is reduced to approximately 25% of in vivo Zn2+ release. While physiological concentration of extracellular Zn2+ is low nanomolar in young rat brain, it may be elevated along with aging (Figure 1), based on the age-related increase in extracellular zinc concentration in the hippocampus [38].

3. Impact of Synaptic Zn2+ Dynamics on Cognition

Extracellular Ca2+ concentration is not affected by neuronal excitation. In contrast, extracellular Zn2+ concentration is dynamically increased by zincergic excitation, but not by non-zincergic excitation. In any case, extracellular dynamics of Ca2+ and Zn2+ is critically linked with their intracellular dynamics. The basal concentration of intracellular (cytosol) Ca2+ is 10–100 nM, while that of intracellular Zn2+ is extremely low and estimated to be less than 1 nM (Figure 1) [39,40]. While intracellular Ca2+ serves as a signaling factor for plastic changes at synapses, intracellular Zn2+ is increased for not only signaling for plastic changes during learning and cognitive activity but also plastic changes in synapse structure [41,42]. The optimal range of intracellular Zn2+ increased during learning and cognitive activity, which is dynamically linked with Zn2+ release at zincergic synapses, remains to be clarified. Even at non-zincergic synapses, postsynaptic intracellular Zn2+ may reach ~1 nM and the increase originates in internal stores/proteins unlike the neurotoxic increase via extracellular Zn2+ influx as described below.
LTP at zincergic mossy fiber-CA3 pyramidal cell synapses is induced by the presynaptic mechanism, in which glutamate release is persistently increased. Mossy fiber LTP induction critically depends on the rise in presynaptic Ca2+ [43,44,45], which activates the calcium-calmodulin-sensitive adenyl cyclase I [46]. Zn2+ released from mossy fibers is immediately retaken up into presynaptic terminals through Ca2+ channels and activates a Src family kinase, which promotes tropomyosin-related kinase B (TrkB) activation. The activation leads to the phosphorylation and activation of phospholipase Cγ1, followed by calcium-calmodulin-sensitive adenyl cyclase I activation. Zn2+ increases presynaptic glutamate release, while it inhibits postsynaptic mechanism of mossy fiber LTP via Zn2+ influx [47,48].
LTP at the Schaffer collateral/commissural-CA1 pyramidal cell synapses depends on the postsynaptic activation of NMDA receptors [49]. NMDA receptor activation increases postsynaptic Ca2+ concentration, which leads to LTP and LTD. NMDA receptors consist of multiple subclasses [50] and the subtypes have different sensitivities to Zn2+, an endogenous blocker [51,52,53,54]. ZnAF-2DA is a useful tool to evaluate the direct involvement of Zn2+ in cognitive function. ZnAF-2DA, a membrane-permeable Zn2+ indicator, is taken up into neurons through the plasma membrane and is hydrolyzed by esterase in the cytosol, resulting in the production of ZnAF-2, which cannot permeate the plasma membrane [55,56]. When ZnAF-2DA is locally injected into the hippocampal CA1, intracellular ZnAF-2 is detected only in the injected area in the CA1 and can block cellular transients in Zn2+ concentration (Kd, 2.7 nM for Zn2+). The concurrent evaluations of in vivo LTP and learning behavior in separated experiments using ZnAF-2DA answer whether the in vivo LTP via intracellular Zn2+ signaling is linked with learning behavior. The influx of extracellular Zn2+ into CA1 pyramidal cells, which is linked with Zn2+ release form the zincergic Schaffer collateral, is required for object recognition memory via in vivo Schaffer collateral LTP [57]. Glutamatergic input to CA1 pyramidal cells via the medial perforant pathway (the temporoammonic pathway) from the entorhinal cortex facilitates memory consolidation [58] and is required for temporal association memory [59] and spatial working memory [60]. Although the medial perforant pathway from the entorhinal cortex, is non-zincergic [61], intracellular Zn2+ signaling, which originates in internal stores/proteins, is required for LTP at medial perforant pathway-CA1 pyramidal cell synapses [62]. It is likely that intracellular Zn2+ signaling in CA1 pyramidal cells is also involved in cognitive function via in vivo perforant pathway LTP.
The lateral and medial entorhinal cortices are connected with the dentate gyrus. The lateral and the medial perforant pathways, which originate in the lateral and the medial entorhinal cortices, respectively, comprise physiologically distinct inputs to the dentate gyrus. The lateral perforant pathway transmits nonspatial information, while the medial perforant pathway transmits spatial information [63]. In regard to LTP at medial perforant pathway-dentate granule cell synapses, calmodulin-dependent protein kinase II α (α-CaMKII)/brain-derived neurotrophic factor (BDNF) signaling pathway plays a key role for LTP induction. Zinc deficiency-induced cognitive and synaptic impairments are linked with disruption of α-CaMKII/BDNF signaling pathway [64]. In dentate granule cells, intracellular Zn2+ signaling originates in internal stores/proteins and is necessary for object and space recognition memory via medial perforant pathway LTP [65,66]. In postsynaptic neurons innervated by non-zincergic medial perforant pathway, glutamate receptor activation triggers off Zn2+ release from internal stores/proteins that remain to be clarified.
Hippocampal neurogenesis always produces dentate granule cells, in which NMDA receptor-dependent synaptic plasticity is involved in learning and memory [67,68]. Zn2+ is concentrated in the dentate gyrus of the hippocampus [69] and is required for neurogenesis process [70]. In human neuronal precursor cells, zinc deficiency induces apoptosis via mitochondrial p53- and caspase-dependent pathways [71], suggesting that dynamic Zn2+ transport to neuronal precursor cells is critical for learning and memory via hippocampal neurogenesis [72].

4. Impact of Synaptic Zn2+ Dynamics on Cognitive Decline

Aging has progressive pathophysiological features and is linked with altered cell metabolism, damaged nucleic acid, oxidative stress, and deposition of abnormal forms of proteins. Aging also is characterized by cognitive decline, neuronal loss, and vulnerability to neurological disorders [73] and may be often related with altered Zn2+ homeostasis in the brain [74,75]. Hippocampal zinc concentration is decreased in aging, which decreases zinc transporter-3 (ZnT3) protein. ZnT3 controls synaptic vesicular Zn2+ levels. Zn2+ release from zincergic neuron terminals, which dynamically modifies the basal concentration of extracellular Zn2+, is decreased in aging [74], while the basal concentration of extracellular Zn2+ may be increased [38], probably as a compensatory mechanism. A negative modulation of extracellular glutamate signaling by extracellular Zn2+ may be involved in cognitive function.
Metal chaperones i.e., clioquinol and PBT2, prevent normal age-related cognitive decline [76,77], suggest that metal chaperones are effective for preventing Zn2+-mediated cognitive decline that is observed in aging and disease. The hippocampus is vulnerable to Zn2+ neurotoxicity [78] and the dentate gyrus is the most vulnerable to aging process [2,79]. Although the vulnerability to aging is poorly understood, it is possible that synaptic Zn2+ dynamics is involved in the vulnerability. New granule cells are continuously produced in the subgranular zone of the dentate gyrus (Figure 2) and the decreased rate of hippocampal neurogenesis is involved in age-related cognitive decline [80]. Neurogenesis-related apoptosis, which seems to be increased along with aging, always occurs in the dentate gyrus. In the subgranular zone, the apoptosis locally increases extracellular K+ and the increase is due to the efflux of intracellular K+ (approximately 140 mM) by disruption of the plasma membrane. The increase in extracellular K+ may excite granule cells and pyramidal basket cells, which exist nearby in the dentate gyrus, and disturbs intracellular dynamics of Ca2+ and Zn2+ (Figure 2). As a matter of fact, both memory acquisition via LTP induction and memory retention via LTP maintenance are impaired after local injection of high K+ into the dentate gyrus [81,82,83] or the CA1 [84]. The impairments are due to an increase in intracellular Zn2+, but not that in intracellular Ca2+, because the impairments are rescued with CaEDTA, which forms membrane-impermeable ZnEDTA in the extracellular compartment and inhibits the influx of extracellular Zn2+, but not that in extracellular Ca2+ [82]. The evidence indicates Zn2+-mediated cognitive decline via transient Zn2+ accumulation in dentate granule cells (Figure 1) and CA1 pyramidal cells.
If the basal level of extracellular Zn2+ is increased age-dependently in the hippocampus (Figure 1) [38], it is estimated that Zn2+-mediated cognitive decline more readily occurs in the aged brain. High K+-induced increase in intracellular Zn2+ is facilitated in the aged dentate gyrus and leads to attenuating both LTP induction and maintained LTP at medial perforant pathway-dentate granule cell synapses of aged rats [38,83], suggesting that the influx of extracellular Zn2+ into dentate granule cells more readily occurs in aged rats and is a cause of age-related cognitive decline via attenuation of LTP. It is likely that neurogenesis-related apoptosis is involved in Zn2+-mediated cognitive decline.
GluR2-lacking calcium-permeable α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) receptors are involved in Zn2+-mediated neurodegeneration in the hippocampal CA1 and CA3 [18,85,86]. In the hippocampus, the levels of GluR1 and GluR2 mRNA are highest in the dentate gyrus and the GluR1/GluR2 mRNA ratios are elevated along with aging [87]. The findings suggest that Zn2+ influx through Ca2+-permeable AMPA receptors, which more readily occurs in the aged dentate gyrus, plays a key role for cognitive decline [38,83]. Intracellular Zn2+ can reach approximately 10 nM via the rapid influx of extracellular Zn2+ (Figure 1) [34]. Both increases in extracellular Zn2+ and Ca2+-permeable AMPA receptors contribute to Zn2+-mediated cognitive decline in aging.
Although intracellular Zn2+ level in the process of LTP maintenance is unknown, LTP maintenance at medial perforant pathway-dentate granule cell synapses is affected by chelation of intracellular Zn2+ with intracellular ZnAF-2 [66] and the aged dentate gyrus is more susceptible to the chelating effect on LTP maintenance [83]. When ZnAF-2DA is used as an index of the capacity binding intracellular Zn2+, interestingly, the capacity of intracellular ZnAF-2 for binding intracellular Zn2+ is more rapidly lost in the aged dentate molecular layer where medial perforant pathway-dentate granule cell synapses are contained than in the young dentate molecular layer, suggesting that intracellular Zn2+-buffering is weakened in the dentate gyrus along with aging (Figure 3) [83]. Characteristics (easiness) of extracellular Zn2+ influx may be linked with weakened intracellular Zn2+-buffering in the aged dentate gyrus [28]. Although the actual state of intracellular Zn2+-buffering is poorly understood, Ca2+-permeable channels, zinc transporters (ZIP and ZnT), zinc-binding proteins such as metallothioneins, and Zn2+-containing internal stores are involved in the Zn2+-buffering system.

5. Perspectives

Vulnerability to Ca2+ dysregulation has been observed in the process of brain aging [88,89,90]. It has been reported that Ca2+ dysregulation is not ubiquitous. The mechanisms of dysregulation are observed in specific cell populations and areas in the brain. For example, L-type Ca2+ channels is age-dependently increased in hippocampal pyramidal cells [91]. Age-dependent reduction in the NMDA receptor function is observed in the hippocampus and the frontal cortex [92], suggesting a compensatory mechanism to availability/restriction for intracellular Ca2+ signaling. Intracellular Ca2+-buffering, which is involved in cognitive function, is weakened during brain aging [89]. In contrast, intracellular Zn2+-buffering is also dynamically involved in cognition and its decline. However, the Zn2+-buffering system is more poorly understood than the Ca2+-buffering system, and its clarification is required for understanding cognition and its decline.

Author Contributions

Atsushi Takeda wrote the manuscript and Haruna Tamano confirmed the contents.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Knierim, J.J. The hippocampus. Curr. Biol. 2015, 25, R1116–R1121. [Google Scholar] [CrossRef] [PubMed]
  2. Small, S.A.; Schobel, S.A.; Buxton, R.B.; Witter, M.P.; Barnes, C.A. A pathophysiological framework of hippocampal dysfunction in ageing and disease. Nat. Rev. Neurosci. 2011, 12, 585–601. [Google Scholar] [CrossRef] [PubMed]
  3. Sasaki, T.; Leutgeb, S.; Leutgeb, J.K. Spatial and memory circuits in the medial entorhinal cortex. Curr. Opin. Neurobiol. 2015, 32, 16–23. [Google Scholar] [CrossRef] [PubMed]
  4. Malenka, R.C.; Nicoll, R.A. Long-term potentiation—A decade of progress? Science 1999, 285, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  5. Blitzer, R.D.; Iyengar, R.; Landau, E.M. Postsynaptic signaling networks: Cellular cogwheels underlying long-term plasticity. Biol. Psychiatry 2005, 57, 113–119. [Google Scholar] [CrossRef] [PubMed]
  6. Nicoll, R.A.; Malenka, R.C. Contrasting properties of two forms of long-term potentiation in the hippocampus. Nature 1995, 377, 115–118. [Google Scholar] [CrossRef] [PubMed]
  7. Danbolt, N.C. Glutamate uptake. Prog. Neurobiol. 2001, 65, 1–105. [Google Scholar] [CrossRef]
  8. Dong, X.X.; Wang, Y.; Qin, Z.H. Molecular mechanisms of excitotoxicity and their relevance to pathogenesis of neurodegenerative diseases. Acta Pharmacol. Sin. 2009, 30, 379–387. [Google Scholar] [CrossRef] [PubMed]
  9. Lai, T.W.; Zhang, S.; Wang, Y.T. Excitotoxicity and stroke: Identifying novel targets for neuroprotection. Prog. Neurobiol. 2014, 115, 157–188. [Google Scholar] [CrossRef] [PubMed]
  10. Lewerenz, J.; Maher, P. Chronic glutamate toxicity in neurodegenerative diseases—What is the evidence? Front. Neurosci. 2015, 9, 469. [Google Scholar] [CrossRef] [PubMed]
  11. Lo, E.H.; Dalkara, T.; Moskowitz, M.A. Mechanisms, challenges and opportunities in stroke. Nat. Rev. Neurosci. 2003, 4, 399–415. [Google Scholar] [CrossRef] [PubMed]
  12. Malairaman, U.; Dandapani, K.; Katyal, A. Effect of Ca2EDTA on zinc mediated inflammation and neuronal apoptosis in hippocampus of an in vivo mouse model of hypobaric hypoxia. PLoS ONE 2014, 9, e110253. [Google Scholar] [CrossRef] [PubMed]
  13. Medvedeva, Y.V.; Weiss, J.H. Intramitochondrial Zn2+ accumulation via the Ca2+ uniporter contributes to acute ischemic neurodegeneration. Neurobiol. Dis. 2014, 68, 137–144. [Google Scholar] [CrossRef] [PubMed]
  14. Tamano, H.; Koike, Y.; Nakada, H.; Shakushi, Y.; Takeda, A. Significance of synaptic Zn2+ signaling in zincergic and non-zincergic synapses in the hippocampus in cognition. J. Trace Elem. Med. Biol. 2016, 38, 93–98. [Google Scholar] [CrossRef] [PubMed]
  15. Takeda, A.; Tamano, H. Significance of the degree of synaptic Zn2+ signaling in cognition. BioMetals 2016, 29, 177–185. [Google Scholar] [CrossRef] [PubMed]
  16. Takeda, A.; Tamano, H. Insight into zinc signaling from dietary zinc deficiency. Brain Res. Rev. 2009, 62, 33–34. [Google Scholar] [CrossRef] [PubMed]
  17. Lavoie, N.; Jeyaraju, D.V.; Peralta, M.R., 3rd; Seress, L.; Pellegrini, L.; Tóth, K. Vesicular zinc regulates the Ca2+ sensitivity of a subpopulation of presynaptic vesicles at hippocampal mossy fiber terminals. J. Neurosci. 2011, 31, 18251–18265. [Google Scholar] [CrossRef] [PubMed]
  18. Colbourne, F.; Grooms, S.Y.; Zukin, R.S.; Buchan, A.M.; Bennett, M.V. Hypothermia rescues hippocampal CA1 neurons and attenuates down-regulation of the AMPA receptor GluR2 subunit after forebrain ischemia. Proc. Natl. Acad. Sci. USA 2003, 100, 2906–2910. [Google Scholar] [CrossRef] [PubMed]
  19. Frederickson, C.J.; Koh, J.Y.; Bush, A.I. The neurobiology of zinc in health and disease. Nat. Rev. Neurosci. 2005, 6, 449–462. [Google Scholar] [CrossRef] [PubMed]
  20. Noh, K.M.; Yokota, H.; Mashiko, T.; Castillo, P.E.; Zukin, R.S.; Bennett, M.V. Blockade of calcium-permeable AMPA receptors protects hippocampal neurons against global ischemia-induced death. Proc. Natl. Acad. Sci. USA 2005, 102, 12230–12235. [Google Scholar] [CrossRef] [PubMed]
  21. Stork, C.J.; Li, Y.V. Rising zinc: A significant cause of ischemic neuronal death in the CA1 region of rat hippocampus. J. Cereb. Blood Flow Metab. 2009, 29, 1399–1408. [Google Scholar] [CrossRef] [PubMed]
  22. Sensi, S.L.; Paoletti, P.; Bush, A.I.; Sekler, I. Zinc in the physiology and pathology of the CNS. Nat. Rev. Neurosci. 2009, 10, 780–791. [Google Scholar] [CrossRef] [PubMed]
  23. Krężel, A.; Maret, W. The biological inorganic chemistry of zinc ions. Arch. Biochem. Biophys. 2016, 611, 3–19. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, N.K.; Robinson, H.P. Effects of divalent cations on slow unblock of native NMDA receptors in mouse neocortical pyramidal neurons. Eur. J. Neurosci. 2011, 34, 199–212. [Google Scholar] [CrossRef] [PubMed]
  25. Jones, H.C.; Keep, R.F. Brain fluid calcium concentration and response to acute hypercalcaemia during development in the rat. J. Physiol. 1988, 402, 579–593. [Google Scholar] [CrossRef] [PubMed]
  26. Neves, G.; Cooke, S.F.; Bliss, T.V.P. Synaptic plasticity, memory and the hippocampus: A neural network approach to causality. Nat. Rev. Neurosci. 2008, 9, 65–67. [Google Scholar] [CrossRef] [PubMed]
  27. Lisman, J.; Yasuda, R.; Raghavachari, S. Mechanisms of CaMKII action in long-term potentiation. Nat. Rev. Neurosci. 2012, 13, 169–182. [Google Scholar] [CrossRef] [PubMed]
  28. Alberdi, E.; Sánchez-Gómez, M.V.; Cavaliere, F.; Pérez-Samartín, A.; Zugaza, J.L.; Trullas, R.; Domercq, M.; Matute, C. Amyloid β oligomers induce Ca2+ dysregulation and neuronal death through activation of ionotropic glutamate receptors. Cell Calcium 2010, 47, 264–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Bickler, P.E.; Warren, D.E.; Clark, J.P.; Gabatto, P.; Gregersen, M.; Brosnan, H. Anesthetic protection of neurons injured by hypothermia and rewarming: Roles of intracellular Ca2+ and excitotoxicity. Anesthesiology 2012, 117, 280–292. [Google Scholar] [CrossRef] [PubMed]
  30. Abushik, P.A.; Sibarov, D.A.; Eaton, M.J.; Skatchkov, S.N.; Antonov, S.M. Kainate-induced calcium overload of cortical neurons in vitro: Dependence on expression of AMPAR GluA2-subunit and down-regulation by subnanomolar ouabain. Cell Calcium 2013, 54, 95–104. [Google Scholar] [CrossRef] [PubMed]
  31. Hershey, C.O.; Hershey, L.A.; Varnes, A.; Vibhakar, S.D.; Lavin, P.; Strain, W.H. Cerebrospinal fluid trace element content in dementia: Clinical, radiologic, and pathologic correlations. Neurology 1983, 33, 1350–1353. [Google Scholar] [CrossRef] [PubMed]
  32. Gellein, K.; Skogholt, J.H.; Aaseth, J.; Thoresen, G.B.; Lierhagen, S.; Steinnes, E.; Syversen, T.; Flaten, T.P. Trace elements in cerebrospinal fluid and blood from patients with a rare progressive central and peripheral demyelinating disease. J. Neurol. Sci. 2008, 266, 70–78. [Google Scholar] [CrossRef] [PubMed]
  33. Michalke, B.; Nischwitz, V. Review on metal speciation analysis in cerebrospinal fluid-current methods and results: A review. Anal. Chim. Acta 2010, 682, 23–36. [Google Scholar] [CrossRef] [PubMed]
  34. Frederickson, C.J.; Giblin, L.J.; Krezel, A.; McAdoo, D.J.; Muelle, R.N.; Zeng, Y.; Balaji, R.V.; Masalha, R.; Thompson, R.B.; Fierke, C.A.; et al. Concentrations of extracellular free zinc (pZn)e in the central nervous system during simple anesthetization, ischemia and reperfusion. Exp. Neurol. 2006, 198, 285–293. [Google Scholar] [CrossRef] [PubMed]
  35. Takeda, A.; Tamano, H. Significance of low nanomolar concentration of Zn2+ in artificial cerebrospinal fluid. Mol. Neurobiol. 2017, 54, 2477–2482. [Google Scholar] [CrossRef] [PubMed]
  36. Tamano, H.; Nishio, R.; Shakushi, Y.; Sasaki, M.; Koike, Y.; Osawa, M.; Takeda, A. In vitro and in vivo physiology of low nanomolar concentrations of Zn2+ in artificial cerebrospinal fluid. Sci. Rep. 2017, 7, 42897. [Google Scholar] [CrossRef] [PubMed]
  37. Suh, S.W.; Danscher, G.; Jensen, M.S.; Thompson, R.; Motamedi, M.; Frederickson, C.J. Release of synaptic zinc is substantially depressed by conventional brain slice preparations. Brain Res. 2000, 879, 7–12. [Google Scholar] [CrossRef]
  38. Takeda, A.; Koike, Y.; Osaw, M.; Tamano, H. Characteristic of extracellular Zn2+ influx in the middle-aged dentate gyrus and its involvement in attenuation of LTP. Mol. Neurobiol. 2017. [Google Scholar] [CrossRef] [PubMed]
  39. Sensi, S.L.; Canzoniero, L.M.T.; Yu, S.P.; Ying, H.S.; Koh, J.Y.; Kerchner, G.A.; Choi, D.W. Measurement of intracellular free zinc in living cortical neurons: Routes of entry. J. Neurosci. 1997, 15, 9554–9564. [Google Scholar]
  40. Colvin, R.A.; Bush, A.I.; Volitakis, I.; Fontaine, C.P.; Thomas, D.; Kikuchi, K.; Holmes, W.R. Insights into Zn2+ homeostasis in neurons from experimental and modeling studies. Am. J. Physiol. Cell Physiol. 2008, 294, C726–C742. [Google Scholar] [CrossRef] [PubMed]
  41. Grabrucker, A.M.; Knight, M.J.; Proepper, C.; Bockmann, J.; Joubert, M.; Rowan, M.; Nienhaus, G.U.; Garner, C.C.; Bowie, J.U.; Kreutz, M.R.; et al. Concerted action of zinc and ProSAP/Shank in synaptogenesis and synapse maturation. EMBO J. 2011, 30, 569–581. [Google Scholar] [CrossRef] [PubMed]
  42. Perrin, L.; Roudeau, S.; Carmona, A.; Domart, F.; Petersen, J.D.; Bohic, S.; Yang, Y.; Cloetens, P.; Ortega, R. Zinc and Copper Effects on Stability of Tubulin and Actin Networks in Dendrites and Spines of Hippocampal Neurons. ACS Chem. Neurosci. 2017, 8, 1490–1499. [Google Scholar] [CrossRef] [PubMed]
  43. Castillo, P.E.; Weisskopf, M.G.; Nicoll, R.A. The role of Ca2+ channels in hippocampal mossy fiber synaptic transmission and long-term potentiation. Neuron 1994, 12, 261–269. [Google Scholar] [CrossRef]
  44. Tong, G.; Malenka, R.C.; Nicoll, R.A. Long-term potentiation in cultures of single hippocampal granule cells: A presynaptic form of plasticity. Neuron 1996, 16, 1147–1157. [Google Scholar]
  45. Breustedt, J.; Vogt, K.E.; Miller, R.J.; Nicoll, R.A.; Schmitz, D. α1E-containing Ca2+ channels are involved in synaptic plasticity. Proc. Natl. Acad. Sci. USA 2003, 100, 12450–12455. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, H.; Storm, D.R. Calmodulin-regulated adenylyl cyclases: Cross-talk and plasticity in the central nervous system. Mol. Pharmacol. 2003, 63, 463–468. [Google Scholar] [CrossRef] [PubMed]
  47. Huang, Y.Z.; Pan, E.; Xiong, Z.Q.; McNamara, J.O. Zinc-mediated transactivation of TrkB potentiates the hippocampal mossy fiber-CA3 pyramid synapse. Neuron 2008, 57, 546–558. [Google Scholar] [CrossRef] [PubMed]
  48. Pan, E.; Zhang, X.A.; Huang, Z.; Krezel, A.; Zhao, M.; Tinberg, C.E.; Lippard, S.J.; McNamara, J.O. Vesicular zinc promotes presynaptic and inhibits postsynaptic long-term potentiation of mossy fiber-CA3 synapse. Neuron 2011, 71, 1116–1126. [Google Scholar] [CrossRef] [PubMed]
  49. Malenka, R.C.; Bear, M.F. LTP and LTD: An embarrassment of riches. Neuron 2004, 44, 5–21. [Google Scholar] [CrossRef] [PubMed]
  50. Cull-Candy, S.G.; Leszkiewicz, D.N. Role of distinct NMDA receptor subtypes at central synapses. Sci. STKE 2004, 255, re16. [Google Scholar] [CrossRef] [PubMed]
  51. Westbrook, G.L.; Mayer, M.L. Micromolar concentrations of Zn2+ antagonize NMDA and GABA responses of hippocampal neurons. Nature 1987, 328, 640–643. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, N.; Moshaver, A.; Raymond, L.A. Differential sensitivity of recombinant N-methyl-d-aspartate receptor subtypes to zinc inhibition. Mol. Pharmacol. 1997, 51, 1015–1023. [Google Scholar] [PubMed]
  53. Paoletti, P.; Ascher, P.; Neyton, J. High-affinity zinc inhibition of NMDA NR1-NR2A receptors. J. Neurosci. 1998, 17, 5711–5725. [Google Scholar]
  54. Choi, Y.B.; Lipton, S.A. Identification and mechanism of action of two histidine residues underlying high-affinity Zn2+ inhibition of the NMDA receptor. Neuron 1999, 23, 171–180. [Google Scholar] [CrossRef]
  55. Hirano, T.; Kikuchi, K.; Urano, Y.; Nagano, T. Improvement and biological applications of fluorescent probes for zinc, ZnAFs. J. Am. Chem. Soc. 2002, 124, 6555–6562. [Google Scholar] [CrossRef] [PubMed]
  56. Ueno, S.; Tsukamoto, M.; Hirano, T.; Kikuchi, K.; Yamada, M.K.; Nishiyama, N.; Nagano, T.; Matsuki, N.; Ikegaya, Y. Mossy fiber Zn2+ spillover modulates heterosynaptic N-methyl-d-aspartate receptor activity in hippocampal CA3 circuits. J. Cell Biol. 2002, 158, 215–220. [Google Scholar] [CrossRef] [PubMed]
  57. Takeda, A.; Suzuki, M.; Tempaku, M.; Ohashi, K.; Tamano, H. Influx of extracellular Zn2+ into the hippocampal CA1 neurons is required for cognitive performance via long-term potentiation. Neuroscience 2015, 304, 209–216. [Google Scholar] [CrossRef] [PubMed]
  58. Remondes, M.; Schuman, E.M. Role for a cortical input to hippocampal area CA1 in the consolidation of a long-term memory. Nature 2004, 431, 699–703. [Google Scholar] [CrossRef] [PubMed]
  59. Suh, J.; Rivest, A.J.; Nakashiba, T.; Tominaga, T.; Tonegawa, S. Entorhinal cortex layer III input to the hippocampus is crucial for temporal association memory. Science 2011, 334, 1415–1420. [Google Scholar] [CrossRef] [PubMed]
  60. Vago, D.R.; Kesner, R.P. Disruption of the direct perforant path input to the CA1 subregion of the dorsal hippocampus interferes with spatial working memory and novelty detection. Behav. Brain Res. 2008, 189, 273–283. [Google Scholar] [CrossRef] [PubMed]
  61. Sindreu, C.B.; Varoqui, H.; Erickson, J.D.; Pérez-Clausell, J. Boutons containing vesicular zinc define a subpopulation of synapses with low AMPAR content in rat hippocampus. Cereb. Cortex 2003, 13, 823–829. [Google Scholar] [CrossRef] [PubMed]
  62. Tamano, H.; Nishio, R.; Takeda, A. Involvement of intracellular Zn2+ signaling in LTP at perforant pathway-CA1 pyramidal cell synapse. Hippocampus 2017, 27, 777–783. [Google Scholar] [CrossRef] [PubMed]
  63. Gonzalez, J.; Morales, I.S.; Villarreal, D.M.; Derrick, B.E. Low-frequency stimulation induces long-term depression and slow onset long-term potentiation at perforant path-dentate gyrus synapses in vivo. J. Neurophysiol. 2014, 111, 1259–1273. [Google Scholar] [CrossRef] [PubMed]
  64. Yu, X.; Ren, T.; Yu, X. Disruption of calmodulin-dependent protein kinase II α/brain-derived neurotrophic factor (α-CaMKII/BDNF) signalling is associated with zinc deficiency-induced impairments in cognitive and synaptic plasticity. Br. J. Nutr. 2013, 110, 2194–2200. [Google Scholar] [CrossRef] [PubMed]
  65. Takeda, A.; Tamano, H.; Ogawa, T.; Takada, S.; Nakamura, M.; Fujii, H.; Ando, M. Intracellular Zn2+ signaling in the dentate gyrus is required for object recognition memory. Hippocampus 2014, 24, 1404–1412. [Google Scholar] [CrossRef] [PubMed]
  66. Tamano, H.; Minamino, T.; Fujii, H.; Takada, S.; Nakamura, M.; Ando, M.; Takeda, A. Blockade of intracellular Zn2+ signaling in the dentate gyrus erases recognition memory via impairment of maintained LTP. Hippocampus 2015, 25, 952–962. [Google Scholar] [CrossRef] [PubMed]
  67. Schmidt-Hieber, C.; Jonas, P.; Bischofberger, J. Enhanced synaptic plasticity in newly generated granule cells of the adult hippocampus. Nature 2004, 429, 184–187. [Google Scholar] [CrossRef] [PubMed]
  68. Kheirbek, M.A.; Tannenholz, L.; Hen, R. NR2B-dependent plasticity of adult-born granule cells is necessary for context discrimination. J. Neurosci. 2012, 32, 8696–8702. [Google Scholar] [CrossRef] [PubMed]
  69. Takeda, A.; Sawashita, J.; Okada, S. Biological half-lives of zinc and manganese in rat brain. Brain Res. 1995, 695, 53–58. [Google Scholar] [CrossRef]
  70. Suh, S.W.; Won, S.J.; Hamby, A.M.; Yoo, B.H.; Fan, Y.; Sheline, C.T.; Tamano, H.; Takeda, A.; Liu, J. Decreased brain zinc availability reduces hippocampal neurogenesis in mice and rats. J. Cereb. Blood Flow Metab. 2009, 29, 1579–1588. [Google Scholar] [CrossRef] [PubMed]
  71. Seth, R.; Corniola, R.S.; Gower-Winter, S.D.; Morgan, T.J., Jr.; Bishop, B.; Levenson, C.W. Zinc deficiency induces apoptosis via mitochondrial p53- and caspase-dependent pathways in human neuronal precursor cells. J. Trace Elem. Med. Biol. 2015, 30, 59–65. [Google Scholar] [CrossRef] [PubMed]
  72. Cope, E.C.; Morris, D.R.; Gower-Winter, S.D.; Brownstein, N.C.; Levenson, C.W. Effect of zinc supplementation on neuronal precursor proliferation in the rat hippocampus after traumatic brain injury. Exp. Neurol. 2016, 279, 96–103. [Google Scholar] [CrossRef] [PubMed]
  73. Mocchegiani, E.; Bertoni-Freddari, C.; Marcellini, F.; Malavolta, M. Brain, aging and neurodegeneration: Role of zinc ion availability. Prog. Neurobiol. 2005, 75, 367–390. [Google Scholar] [CrossRef] [PubMed]
  74. Adlard, P.A.; Parncutt, J.M.; Finkelstein, D.I.; Bush, A.I. Cognitive loss in zinc transporter-3 knock-out mice: A phenocopy for the synaptic and memory deficits of Alzheimer’s disease? J. Neurosci. 2010, 30, 1631–1636. [Google Scholar] [CrossRef] [PubMed]
  75. Sindreu, C.; Palmiter, R.D.; Storm, D.R. Zinc transporter ZnT-3 regulates presynaptic Erk1/2 signaling and hippocampus-dependent memory. Proc. Natl. Acad. Sci. USA 2012, 108, 3366–3370. [Google Scholar] [CrossRef] [PubMed]
  76. Adlard, P.A.; Sedjahtera, A.; Gunawan, L.; Bray, L.; Hare, D.; Lear, J.; Doble, P.; Bush, A.I.; Finkelstein, D.I.; Cherny, R.D. A novel approach to rapidly prevent age-related cognitive decline. Aging Cell 2014, 13, 351–359. [Google Scholar] [CrossRef] [PubMed]
  77. Adlard, P.A.; Parncutt, J.; Lal, V.; James, S.; Hare, D.; Doble, P.; Finkelstein, D.I.; Bush, A.I. Metal chaperones prevent zinc-mediated cognitive decline. Neurobiol. Dis. 2015, 81, 196–202. [Google Scholar] [CrossRef] [PubMed]
  78. Stork, C.J.; Li, Y.V. Intracellular zinc elevation measured with a “calcium-specific” indicator during ischemia and reperfusion in rat hippocampus: A question on calcium overload. J. Neurosci. 2006, 26, 10430–10437. [Google Scholar] [CrossRef] [PubMed]
  79. Brickman, A.M.; Khan, U.A.; Provenzano, F.A.; Yeung, L.K.; Suzuki, W.; Schroeter, H.; Wall, M.; Sloan, R.P.; Small, S.A. Enhancing dentate gyrus function with dietary flavanols improves cognition in older adults. Nat. Neurosci. 2014, 17, 1798–1803. [Google Scholar] [CrossRef] [PubMed]
  80. Jinno, S. Aging affects new cell production in the adult hippocampus: A quantitative anatomic review. J. Chem. Neuroanat. 2016, 76, 64–72. [Google Scholar] [CrossRef] [PubMed]
  81. Suzuki, M.; Fujise, Y.; Tsuchiya, Y.; Tamano, H.; Takeda, A. Excess influx of Zn2+ into dentate granule cells affects object recognition memory via attenuated LTP. Neurochem. Int. 2015, 87, 60–65. [Google Scholar] [CrossRef] [PubMed]
  82. Takeda, A.; Tamano, H.; Hisatsune, M.; Murakami, T.; Nakada, H.; Fujii, H. Maintained LTP and memory are lost by Zn2+ influx into dentate granule cells, but not Ca2+ influx. Mol. Neurobiol. 2017. [Google Scholar] [CrossRef] [PubMed]
  83. Takeda, A.; Tamano, H.; Murakami, T.; Nakada, H.; Minamino, T.; Koike, Y. Weakened intracellular Zn2+-buffering in the aged dentate gyrus and its involvement in erasure of maintained LTP. Mol. Neurobiol. 2017. [Google Scholar] [CrossRef] [PubMed]
  84. Takeda, A.; Takada, S.; Nakamura, M.; Suzuki, M.; Tamano, H.; Ando, M.; Oku, N. Transient increase in Zn2+ in hippocampal CA1 pyramidal neurons causes reversible memory deficit. PLoS ONE 2011, 6, e28615. [Google Scholar] [CrossRef] [PubMed]
  85. Liu, S.; Lau, L.; Wei, J.; Zhu, D.; Zou, S.; Sun, H.S.; Fu, Y.; Liu, F.; Lu, Y. Expression of Ca(2+)-permeable AMPA receptor channels primes cell death in transient forebrain ischemia. Neuron 2004, 43, 43–55. [Google Scholar] [CrossRef] [PubMed]
  86. Weiss, J.H. Ca permeable AMPA channels in diseases of the nervous system. Front. Mol. Neurosci. 2011, 4, 42. [Google Scholar] [CrossRef] [PubMed]
  87. Pagliusi, S.R.; Gerrard, P.; Abdallah, M.; Talabot, D.; Catsicas, S. Age-related changes in expression of AMPA-selective glutamate receptor subunits: Is calcium-permeability altered in hippocampal neurons? Neuroscience 1994, 61, 429–433. [Google Scholar] [CrossRef]
  88. Foster, T.C. Calcium homeostasis and modulation of synaptic plasticity in the aged brain. Aging Cell 2007, 6, 319–325. [Google Scholar] [CrossRef] [PubMed]
  89. Kumar, A.; Bodhinathan, K.; Foster, T.C. Susceptibility to calcium dysregulation during brain aging. Front. Aging Neurosci. 2009, 1, 2. [Google Scholar] [CrossRef] [PubMed]
  90. Toescu, E.C.; Vreugdenhil, M. Calcium and normal brain ageing. Cell Calcium 2010, 47, 158–164. [Google Scholar] [CrossRef] [PubMed]
  91. Thibault, O.; Landfield, P.W. Increase in single L-type calcium channels in hippocampal neurons during aging. Science 1996, 272, 1017–1020. [Google Scholar] [CrossRef] [PubMed]
  92. Berridge, M.J. Dysregulation of neural calcium signaling in Alzheimer disease, bipolar disorder and schizophrenia. Prion 2013, 7, 2–13. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Zn2+-mediated cognitive decline via rapid influx of extracellular Zn2+. Estimated basal concentration of extracellular Zn2+ is ~10 nM in the adult brain. When the basal concentration of extracellular Zn2+ reaches 100 nM in the adult brain, it induces cognitive decline. In contrast, even if the basal concentration of extracellular Zn2+ reaches 100 nM in the aged brain, it does not induce cognitive decline, suggesting that the basal concentration of extracellular Zn2+ is ~100 nM in the aged brain. Extracellular Zn2+-mediated cognitive decline is induced by glutamateric synapse excitation, in which intracellular Zn2+ may reach 1~10 nM. The synapses are non-zincergic. CC: Ca2+-permeable channels.
Figure 1. Zn2+-mediated cognitive decline via rapid influx of extracellular Zn2+. Estimated basal concentration of extracellular Zn2+ is ~10 nM in the adult brain. When the basal concentration of extracellular Zn2+ reaches 100 nM in the adult brain, it induces cognitive decline. In contrast, even if the basal concentration of extracellular Zn2+ reaches 100 nM in the aged brain, it does not induce cognitive decline, suggesting that the basal concentration of extracellular Zn2+ is ~100 nM in the aged brain. Extracellular Zn2+-mediated cognitive decline is induced by glutamateric synapse excitation, in which intracellular Zn2+ may reach 1~10 nM. The synapses are non-zincergic. CC: Ca2+-permeable channels.
Ijms 18 02411 g001
Figure 2. Neuronal depolarization via neurogenesis-related apoptosis. Neurogenesis-related apoptosis increases extracellular K+ concentration ([K+]o), which is due to the efflux of intracellular K+ as shown by red arrows, in the dentate granule cell layer and can lead dentate granule cells to depolarization, followed by extracellular Zn2+ influx-mediated cognitive decline. The blue arrow shows the process of neurogenesis and red up-arrows show the process of apoptosis and efflux of intracellular K+. [K+]i: intracellular K+ concentration.
Figure 2. Neuronal depolarization via neurogenesis-related apoptosis. Neurogenesis-related apoptosis increases extracellular K+ concentration ([K+]o), which is due to the efflux of intracellular K+ as shown by red arrows, in the dentate granule cell layer and can lead dentate granule cells to depolarization, followed by extracellular Zn2+ influx-mediated cognitive decline. The blue arrow shows the process of neurogenesis and red up-arrows show the process of apoptosis and efflux of intracellular K+. [K+]i: intracellular K+ concentration.
Ijms 18 02411 g002
Figure 3. Is intracellular Zn2+-buffering weakened in the dentate gyrus along with aging? In vivo intracellular Zn2+-buffering is assessed in the dentate molecular layer where non-zincergic media performant pathway (MPP)-dentate granule cell synapses are contained. Intracellular ZnAF-2, an index of intracellular Zn2+-buffering capacity, can bind Zn2+ at young MPP synapse 2 h after ZnAF-2DA injection into the dentate molecular layer, but not at aged MPP synapses. Capacity of intracellular ZnAF-2 for binding intracellular Zn2+ is more rapidly lost in in aged dentate gyrus, probably due to easiness of extracellular Zn2+ influx, suggesting a reduced capacity of intracellular Zn2+-buffering in aged dentate gyrus. ZIP: Zrt/Irt-like proteins.
Figure 3. Is intracellular Zn2+-buffering weakened in the dentate gyrus along with aging? In vivo intracellular Zn2+-buffering is assessed in the dentate molecular layer where non-zincergic media performant pathway (MPP)-dentate granule cell synapses are contained. Intracellular ZnAF-2, an index of intracellular Zn2+-buffering capacity, can bind Zn2+ at young MPP synapse 2 h after ZnAF-2DA injection into the dentate molecular layer, but not at aged MPP synapses. Capacity of intracellular ZnAF-2 for binding intracellular Zn2+ is more rapidly lost in in aged dentate gyrus, probably due to easiness of extracellular Zn2+ influx, suggesting a reduced capacity of intracellular Zn2+-buffering in aged dentate gyrus. ZIP: Zrt/Irt-like proteins.
Ijms 18 02411 g003

Share and Cite

MDPI and ACS Style

Takeda, A.; Tamano, H. The Impact of Synaptic Zn2+ Dynamics on Cognition and Its Decline. Int. J. Mol. Sci. 2017, 18, 2411. https://doi.org/10.3390/ijms18112411

AMA Style

Takeda A, Tamano H. The Impact of Synaptic Zn2+ Dynamics on Cognition and Its Decline. International Journal of Molecular Sciences. 2017; 18(11):2411. https://doi.org/10.3390/ijms18112411

Chicago/Turabian Style

Takeda, Atsushi, and Hanuna Tamano. 2017. "The Impact of Synaptic Zn2+ Dynamics on Cognition and Its Decline" International Journal of Molecular Sciences 18, no. 11: 2411. https://doi.org/10.3390/ijms18112411

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop