Next Article in Journal
Effects of the Acetyltransferase p300 on Tumour Regulation from the Novel Perspective of Posttranslational Protein Modification
Previous Article in Journal
GABRG2 Variants Associated with Febrile Seizures
Previous Article in Special Issue
The Role of Zinc in Modulating Acid-Sensing Ion Channel Function
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Zinc Homeostasis: An Emerging Therapeutic Target for Neuroinflammation Related Diseases

1
College of Pharmacy, Guilin Medical College, Guilin 541199, China
2
Guangxi Key Laboratory of Brain and Cognitive Neuroscience, Guilin Medical College, Guilin 541199, China
3
Basic Medical College, Guilin Medical College, Guilin 541199, China
4
Department of Radiology, Fifth Clinical Medical College, Guilin Medical College, Guilin 541002, China
5
Central Laboratory, Guangxi Health Commission Key Laboratory of Glucose and Lipid Metabolism Disorders, Second Affiliated Hospital of Guilin Medical College, Guilin 541199, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2023, 13(3), 416; https://doi.org/10.3390/biom13030416
Submission received: 28 November 2022 / Revised: 19 December 2022 / Accepted: 20 January 2023 / Published: 22 February 2023
(This article belongs to the Special Issue Zinc in Health and Disease Conditions)

Abstract

:
Zinc is an indispensable trace element in the human body and plays an important role in regulating normal growth and development. Zinc homeostasis in the central nervous system is closely related to the development of neuroinflammation, and synaptic zinc homeostasis disorders affect zinc homeostasis in the brain. Under the condition of synaptic zinc homeostasis, proper zinc supplementation improves the body’s immunity and inhibits neuroinflammation. Synaptic zinc homeostasis disorder in the brain promotes the occurrence and development of neuroinflammation. Cerebral ischemia and hypoxia cause a massive release of synaptic Zn2+ into the synaptic cleft, resulting in neurotoxicity and neuroinflammation. Synaptic zinc homeostasis disorder is a high-risk factor for neurodegenerative diseases. Maintaining cerebral zinc homeostasis suppresses the progression of neuroinflammation-mediated neurodegenerative diseases. This article reviews the relationship between brain zinc homeostasis and neuroinflammation and proposes that maintaining synaptic zinc homeostasis prevents neuroinflammation.

1. Introduction

Neuroinflammation is an inflammatory response that occurs in the central nervous system (CNS). Because the blood-brain barrier (BBB) prevents a large number of peripheral immune cells from entering the brain, neuroinflammation is different from peripheral inflammation, and microglia are the main functional cells in the CNS involved in the inflammatory response. Resting-state microglia are highly branched and monitor the state of other cells in the brain [1]. Under stimulation, resting microglia become polarized to form an amoeba-like shape and become activated [1]. Proinflammatory stimulation induces the polarization of microglia to the M1 type, which induces microglial expression and release of interleukin (IL)-1β, IL-6, and tumor necrosis factor-α (TNF-α) [2]. Anti-inflammatory stimulation induces microglial polarization to the M2 type, which induces microglia to express and release IL-4, IL-10 and transforms growth factor-β [2].
Zinc is the second most abundant trace element in the human body, and it exists stably in the form of Zn2+. Zinc promotes neurogenesis [3], improves immunity [4], acts as a cofactor for enzymes [5], and participates in synaptic transmission [6] and cell apoptosis [7]. Many proteins in the human body, such as the zinc finger protein, contain Zn2+; many enzymes in the body need to combine with Zn2+ to play a catalytic role [5]. Therefore, maintaining zinc homeostasis is important for the normal functioning of the human body, and zinc homeostasis disorder may be one of the causes of many diseases, such as cerebral ischemia and hypoxia [8], epilepsy [9], depression [10], and neurodegenerative diseases [11]. However, the mechanism by which zinc homeostasis disorders cause these diseases needs to be further researched.
Synaptic zinc homeostasis disorder affects brain zinc homeostasis. When synaptic zinc is in a steady state, appropriate supplement of zinc improves the body’s immunity and inhibits neuroinflammation [12,13]. When synaptic zinc is disordered, abnormal release of synaptic Zn2+ leads to severe neuroinflammation. Severe stresses, such as cerebral ischemia and hypoxia [8], and epilepsy [14], lead to a large amount of synaptic Zn2+ being released into the synaptic cleft, which mediates neuronal death and severe neuroinflammation. Neurodegenerative diseases are chronic and progressive diseases with widespread neurotransmitter imbalance. Therefore, maintaining synaptic zinc homeostasis plays an important role in delaying the progression of neurodegenerative diseases [11,15,16]. Neurotoxicity and neuroinflammation caused by disorders of synaptic zinc homeostasis promote the progression of neurodegenerative diseases. This article outlines the relationship between brain zinc homeostasis disorder and neuroinflammation and suggests that maintaining synaptic zinc homeostasis prevents neuroinflammation.

2. Zinc Homeostasis and Its Regulatory Mechanism

2.1. Body Zinc Homeostasis

Zinc homeostasis in the body is regulated by the gastrointestinal tract by absorption of dietary zinc or excretion of zinc by the kidney based on existing levels [17]. More than 90% of the zinc in the body is stored in the skeletal muscle and skeleton, and less than 3% of the total zinc is found in the brain [17]. Zinc in the body exists stably in the form of Zn2+, with one part of the Zn2+ being in the binding state and the other part being in the free state. Zinc in cells is divided into three parts: 1. Zn2+ combined with proteins, including metallothionein (MT); 2. Zn2+ in organelles and vesicles; and 3. Free Zn2+ in the cytoplasm [15,18]. Zrt/Irt-like protein (ZIP) (encoded by the SLC39A gene family) has 14 types of zinc transporters. ZIP transports extracellular Zn2+ as well as Zn2+ in organelles and vesicles to the cytoplasm [17]. ZIP plays an important role in maintaining intracellular zinc homeostasis, and the zinc transporter (ZnT) (encoded by the SLC30A gene family) has 10 ZnT members, among which ZnT9 does not possess the function of zinc transport; hence, only nine types of ZnT transport zinc [17]. The outflow of Zn2+ from the cytoplasm to the extracellular space is mainly mediated by ZnT1, while other zinc transporters mediate the transport of Zn2+ from the cytoplasm to organelles or vesicles [17]. ZnT in the cell membrane is mainly distributed in the organelle membrane, which mediates the uptake of Zn2+ by Golgi bodies, lysosomes, and vesicles [17]. MT is a protein rich in cysteine, which scavenges free radicals [19]. MT reduces the free zinc content in cells by combining with Zn2+ in the cytoplasm and also releases Zn2+ when needed [19]. Therefore, MT is a key protein that regulates cellular zinc homeostasis.

2.2. Brain Zinc Homeostasis and Synaptic Zinc

Zinc in the brain is mainly located in glutamate neurons, most of which are highly enriched in Zn2+ [6]. Highly enriched Zn2+ in the nerve endings of glutamate neurons is released into the synaptic cleft when nerve impulses arrive and enter the postsynaptic membrane, thus regulating the physiological function of the brain [6]. Therefore, an imbalance in synaptic zinc homeostasis significantly affects zinc homeostasis in the brain. Zn2+ binds to N-methyl-d-aspartate (NMDA) and γ-aminobutyric acid type A (GABA-A) receptors, thereby affecting excitatory and inhibitory synaptic transmission in the brain [6]. Synaptic transport of Zn2+ involves the release of Zn2+ from the presynaptic membrane to the postsynaptic membrane and into postsynaptic neurons.
ZnT3 mediates the transfer of Zn2+ from cells to synaptic vesicles [20,21], and synaptic vesicles containing Zn2+ combine with the presynaptic membrane to release Zn2+ into the synaptic cleft. The pathways of Zn2+ entering the postsynaptic neurons in the synaptic cleft include the following [6]: 1. ZIP-mediated transport; 2. Voltage-gated calcium channel (VGCC)-mediated transport; 3. Ca2+-and Zn2+-permeable GluR2-lacking AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid) receptors (AMPARCa-Zn)-mediated transport; 4. NMDA receptor-mediated transport; and 5. Na+/Zn2+ exchanger-mediated transport [22]. The main transport protein of Zn2+ efflux is ZnT1, whereas MT stores Zn2+ and dynamically regulates zinc homeostasis in cells. The release of zinc from presynaptic to postsynaptic neurons increases the level of oxidative stress in postsynaptic neurons, leading to neuronal death [16,23] (as shown in Figure 1).

3. Matrix Metalloproteinases and Neuroinflammation

The catalytically active region of matrix metalloproteinase (MMP) contains a zinc-binding site, which is a zinc-dependent endopeptidase. Under the regulation of inflammatory cytokines and reactive oxygen species (ROS), neurons, astrocytes, and microglia express and release MMP [24,25]. MMP’s function has two sides. MMP promotes tissue remodeling and repair. However, excessive activation also leads to tissue damage. Furthermore, overactivation of MMP destroys the BBB, leading to peripheral immune cells entering the brain and participating in the inflammatory response in the CNS. There is an interaction between neuroinflammation and MMP. IL-1β induces pericyte expression and secretion of MMP-9 by activating the nuclear factor-κB (NF-κB) signaling pathway, and high expression of MMP-9 damages the integrity of the BBB [26]. An in vitro study in which neurons and astrocytes were isolated from mice with Alzheimer’s disease (AD) showed that knockout of the MMP-24 gene significantly reduced IL-1β and β-amyloid (Aβ) in cells [27]. Lipopolysaccharide (LPS) activates the NF-κB signaling pathway by increasing ROS content in astrocytes, thereby upregulating MMP-9 expression [28].
Zinc promotes MMP-mediated neuroinflammation. MMP activation requires Zn2+ binding to the catalytically active region of MMP. Treatment with the Zn2+ chelator clioquinol (CQ) significantly inhibited MMP-2 and MMP-9 activity in hippocampal neurons, indicating that Zn2+ is critical for MMP activation [29]. Treatment of mice with minocycline, an inhibitor of microglial activation, inhibits rotenone-induced MMP-2 and MMP-9 activation, indicating that microglial activation is a key factor mediating MMP activation [30]. Zinc chloride (ZnCl2) treatment enhanced the effect of LPS-induced microglia to release proinflammatory factors, but ROS scavenger inhibited the effect of ZnCl2, suggesting that Zn2+ promoted microglial expression and release of proinflammatory factors by mediating ROS generation [31]. Zn2+ increased the intracellular ROS content by activating reduced nicotinamide adenine dinucleotide phosphate oxidase (NOX) in microglia [8,32]. Zn2+ treatment significantly increased the content of ROS in neurons [33] and enhanced the effect of LPS-induced ROS production in astrocytes [34]. Persistently high concentrations of intracellular and extracellular Zn2+ would promote increased production of ROS and proinflammatory factors, especially in microglia, which mediate immune responses [32,35,36]. The generated ROS and proinflammatory factors increase the activity and expression of MMP, which further promotes neuroinflammation-mediated CNS injury (Figure 1).

4. Zinc Homeostasis and Neuroinflammation

4.1. Zinc Homeostasis Has a Role in Regulating Neuroinflammatory Responses

Zinc sulfate (ZnSO4) downregulated the expression of inducible NO synthase (iNOS), TNF-α, and IL-6 induced by LPS (100 ng/mL) and significantly reduced the content of ROS in cells [37]. A20 is a negative regulator of NF-κB signaling and contains a zinc-finger domain. ZnSO4 inhibits NF-κB signaling by upregulating A20 expression in BV2 cells, thereby inhibiting LPS-induced neuroinflammation [37]. Arginase 1 (Arg-1) is a marker of M2 polarization in microglia. IL-4 induces microglia to express Arg-1 mRNA and increases the cellular zinc in cells [38]. The intracellular Zn2+ chelator TPEN upregulated the expression of Arg-1 mRNA in microglia induced by IL-4, indicating that intracellular free Zn2+ negatively regulates the expression of Arg-1 [38]. In summary, intracellular Zn2+, in cases of zinc homeostasis, inhibits microglial polarization to the M2 type. Under conditions of zinc homeostasis, Zn2+ not only inhibited the expression of proinflammatory factors in microglia but also inhibited the expression of anti-inflammatory factors, playing a role in regulating the neuroinflammatory response.
Zinc supplementation inhibited the activation of microglia in the cerebral cortex of rats fed a high sugar and fat diet, and downregulated the toll-like receptor 4 expression in the rat hippocampus. A high sugar and fat diet damaged the cognitive function of rats, and zinc supplementation significantly improved their cognitive function [4]. Therefore, zinc supplementation improved cognitive function in rats by reducing neuroinflammation. A high-fat diet induced high expression of IL-1β in the cerebral cortex and hippocampus of rats, while zinc supplementation significantly reduced the expression of IL-1β and enhanced novel object recognition in the rats [4]. Thus, zinc supplementation inhibited high-fat diet induced neuroinflammation. Maternal zinc supplementation downregulated the expression of the proinflammatory factors IL-6, IL-1β, and TNF-α in the prefrontal cortex (PFC) of LPS-exposed adult male offspring and decreased the expression of NF-κB in the PFC [13].

4.2. Zinc Homeostasis Imbalance

4.2.1. Low Zinc Promotes Neuroinflammation

TPEN significantly upregulated IL-1β expression in macrophages treated with LPS (1 μg/mL), whereas ZnCl2 treatment reversed the effects of TPEN [39]. TPEN enhanced the release of IL-1β from macrophages induced by LPS [40]. Compared to monocytes cultured at normal Zn2+ concentrations, zinc-deficient monocytes treated with LPS showed significantly upregulated TNF-α and IL-1β expressions, indicating that zinc-deficient monocytes were more sensitive to LPS-induced inflammation [41]. A low-zinc diet or inhibition of zinc absorption changed the intestinal microbiota composition and intestinal function of pregnant mice, upregulating the expression of IL-6 in the brain and activating astrocytes in the hippocampus [42]. Therefore, zinc deficiency may contribute to neuroinflammation by altering intestinal function and intestinal microbiota composition. Prenatal zinc-deficient mice exhibit autism spectrum disorder. A study found that the brains of prenatal zinc-deficient mice showed a neuroinflammatory response, suggesting that prenatal zinc deficiency-mediated neuroinflammation may be the cause of autism in mice [43].

4.2.2. Zinc Accumulation Accelerates Neuroinflammation

Zn2+ increased the content of ROS in microglia by activating NOX in microglia; then, ROS activated the extracellular signal-regulated kinase (ERK) signaling pathway to upregulate the expression of transient receptor potential melastatin 2 (TRPM2) on the cell membrane, which mediates Ca2+ influx [32]. Therefore, microglia were overloaded with Ca2+, leading to cell death. Zeolite-based nanomaterials (Zbn) adsorbed excessive Zn2+ and cleared ROS in the brain of rats after cerebral ischemia, inhibiting the activation of microglia [44]. Therefore, Zbn provided neuroprotection by reducing the content of Zn2+ in the brain. TPEN chelation of Zn2+ inhibited the activation of microglia induced by ZnCl2, indicating that Zn2+ activated microglia and promoted neuroinflammation [35]. Hypoosmotic stress increased the level of extracellular free Zn2+ by upregulating the expression of ZnT1 in astrocytes. Microglia were activated by culture medium obtained by treating astrocytes with hypoosmotic stress, whereas extracellular Zn2+ chelator CaEDTA inhibited microglial activation [45]. Therefore, hypoosmotic stress promoted the release of Zn2+ from astrocytes by upregulating the expression of ZnT1 in astrocytes, thus activating microglia. Zn2+ promoted LPS-induced production of ROS in astrocytes. Zn2+ significantly increased NO production in astrocytes induced by LPS (1 μg/mL) through the p38-mitogen-activated protein kinase signaling pathway [34]. Zn2+ treatment reduced the levels of glutathione (GSH), superoxide dismutase (SOD), and catalase (CAT) in neurons; thus, leading to an increased level of oxidative stress [46]. In conclusion, Zn2+ promoted oxidative stress and neuroinflammation (Figure 1).

4.3. Severe Inflammatory Stress Leads to Zinc Homeostasis Imbalance

Intraperitoneal injection of N-acetyl-L-cysteine (NAC) in rats significantly reduced neuronal death and zinc accumulation induced by cerebral ischemia. NAC downregulated the expression of the non-selective cationic channel TRPM2 in neurons, thereby inhibiting the activation of astrocytes and microglia [47]. Therefore, NAC may reduce the entry of Zn2+ into neurons by downregulating the expression of TRPM2, thus reducing the neuroinflammation mediated by Zn2+. Another study showed that treatment with interferon γ (INFγ) increased Zn2+ content in neurons and upregulated IL-1β, IL-6, and TNF-α expressions in microglia, whereas TRPM2 channel inhibitors reversed the effects of INF γ, indicating that INFγ increased intracellular Zn2+ content by activating TRPM2 channels and promoted neuroinflammation [48]. Therefore, severe inflammatory stress led to an excessive influx of Zn2+ into cells by upregulating the expression of TRPM2, leading to neuroinflammation.

5. Zinc Homeostasis and Neuroinflammatory Diseases

5.1. Zinc Supplementation Promotes the Recovery of Spinal Cord Injury (SCI)

Intraperitoneal injection of zinc gluconate (ZnG) in mice upregulated the expression of Bcl-2 and downregulated the expression of Bax in the injured spinal cord, indicating that ZnG inhibited cell apoptosis in the injured spinal cord [49]. TPEN chelating Zn2+ promoted apoptosis of hippocampal neurons, ERK inhibitor significantly increased TPEN-induced apoptosis of neurons, and ERK agonist reversed the effect of TPEN [50]. Therefore, Zn2+ reduced neuronal apoptosis by activating the ERK signaling pathway. ZnG promoted macrophages to express and release granulocyte colony stimulating factor (G-CSF) [7]. Intraperitoneal injection of ZnG promoted the production of G-CSF in the spinal cord tissue of spinal cord injury (SCI) mice and inhibited the apoptosis of nerve cells, but G-CSF neutralizing antibody reversed this effect [7]. Therefore, ZnG inhibited neuronal apoptosis in the injured spinal cord by promoting G-CSF expression. Treatment with methylprednisolone (MP) reduced neuronal death in rats with SCI, downregulated iNOS, IL-6, and IL-10 expressions in the spinal cord, and inhibited microglial activation. In vitro experiments showed that MP treatment significantly increased Zn2+ and ZIP8 levels in microglia and reversed LPS-induced NF-κB expression in microglia [51]. Therefore, MP may inhibit the NF-κB pathway by promoting the accumulation of Zn2+ in microglia, reducing the release of inflammatory factors and subsequently reducing SCI.
Intraperitoneal injection of ZnG into SCI mice downregulated the expression of NOD-like receptor protein 3 (NLRP3), apoptosis-associated speck-like protein containing a CARD (ASC), caspase-1, and IL-1β in the spinal cord, indicating that zinc treatment inhibited the NLRP3 inflammatory pathway. An in vitro study showed that ZnG inhibited the activation of the LPS-induced NLRP3 inflammatory pathway in microglia [49]. Therefore, Zn2+ reduced the expression and release of inflammatory factors by inhibiting the NLRP3 inflammatory pathway in microglia and accelerating the repair of spinal cord tissues. Intraperitoneal injection of ZnG into SCI mice increased the content of Beclin-1 and the ratio of microtubule-associated protein 1 light chain 3-II/I and reduced the expression of P62. Therefore, ZnG promoted autophagy in nerve cells in the injured spinal cord [49]. The expressions of IL-1β, IL-6, and the NLRP3 inflammasome upregulated in the injured spinal cord, and intraperitoneal injection of ZnG promoted autophagy and inhibited the expression of the NLRP3 inflammasome in SCI mice [52]. An in vitro study showed that ZnG promoted autophagy of microglia and inhibited the expression of the NLRP3 inflammasome, and an autophagy inhibitor reversed the effect of zinc on inhibiting the expression of the NLRP3 inflammasome in microglia. A ubiquitin inhibitor reversed the effect of zinc in promoting NLRP3 inflammasome degradation in microglia, suggesting that zinc promotes NLRP3 degradation through ubiquitination modification [52]. In conclusion, zinc reduced NLRP3 inflammasome content through autophagy and ubiquitination, subsequently promoting repair in SCI mice (Figure 2).

5.2. Zinc Supplement Has Antidepressant Effect

Clinical research has shown that the decrease in zinc concentration in peripheral blood is related to the occurrence and development of depression [53], and zinc supplementation enhances the therapeutic effect of antidepressants [10,54,55]. A low-zinc diet significantly increased the immobility time of mice in the forced swim test [56] and the tail suspension test [57], and decreased the preference for sugar and water [58]. Therefore, the reduction of zinc intake in the body could lead to depression. The combination of ZnCl2 and the natural peptide Cyclo-(His Pro) generated zinc plus Cyclo-(His-Pro) (ZC). Feeding rats with ZC enhanced the body’s absorption of zinc and significantly increased the level of vesicular zinc and neurogenesis in the hippocampus [59]. ZnCl2 treatment upregulated the expression of nestin in induced pluripotent stem cells (iPSCs), whereas TPEN treatment inhibited the expression of nestin in iPSCs, indicating that Zn2+ promoted the differentiation of iPSCs. ZnCl2 treatment promoted ERK and signal transducer and activator of transcription (STAT) phosphorylation, but ERK inhibitors inhibited STAT3 phosphorylation and downregulated nestin expression, indicating that Zn2+ promoted iPSCs differentiation through the ERK-STAT3 signal pathway [60]. Patients with depression show decreased hippocampal neurogenesis, and zinc supplementation promotes hippocampal neurogenesis and has anti-depressive effects. Therefore, zinc supplementation may play an antidepressant role by enhancing neurogenesis.

5.3. Zinc Accumulation Promotes the Development of Acute Cerebral Ischemia

After global cerebral ischemia, the release of Zn2+ from the presynaptic membrane to the postsynaptic membrane led to increased Zn2+ content in postsynaptic neurons, followed by neuronal degeneration, while Zn2+ chelation reduced neuronal degeneration, indicating that the massive release of Zn2+ from the presynaptic membrane is neurotoxic after global cerebral ischemia [61]. Hypobaric hypoxia led to a protein imbalance related to zinc homeostasis, upregulated the expression of TNF-α and iNOS, and also led to neuronal apoptosis in the hippocampus. Zn2+ chelators significantly reversed these effects, indicating that zinc promoted neuroinflammation and apoptosis during hypobaric hypoxia [62]. Chelating Zn2+ by intravenous etidronate-zinc complex (Eti-Zn) inhibited the activation of microglia and decreased the expressions of IL-1β, IL-6, and TNF-α after cerebral ischemia and hypoxia. Eti–Zn also activated the NF-κB signaling pathway, suggesting that Eti–Zn reduced microglial activation by inhibiting the NF-κB signaling pathway, thereby inhibiting neuroinflammation [63]. ZnCl2 significantly induced ROS production in microglia and also enhanced the effects of LPS on the expression of the proinflammatory factors IL-1β, IL-6, and TNF-α [8]. Cerebral ischemia induced microglia to transform into the M1 type and upregulated proinflammatory factor expression in the mouse hippocampus, whereas CaEDTA inhibited its expression in mice after cerebral ischemia, suggesting that chelating Zn2+ attenuated post-ischemic brain injury by inhibiting neuroinflammation [8]. Pretreatment with a high concentration of zinc enhanced the effect of LPS-induced proinflammatory factor secretion in microglia, but this effect was reversed by TPEN and ROS scavengers [31]. Preinjection of CaEDTA into the lateral ventricle inhibited M1 type polarization of microglia in the hippocampus, downregulated the expression of IL-1β, IL-6, and TNF-α, and improved cognitive function in mice after ischemia and reperfusion [31]. In conclusion, zinc accumulation promoted transformation of microglia into a proinflammatory M1 type via ROS, thereby impairing cognitive function in mice after acute cerebral ischemia (Figure 2).

5.4. Zinc Homeostasis Disorder Promotes the Development of Epilepsy

Synaptic zinc homeostasis disorder increased susceptibility to epilepsy [9,64]. Synaptic zinc homeostasis disorder led to an imbalance in central excitation and inhibition, which causes abnormal nerve conduction and triggers epilepsy [9,65]. Knockout of MT3 in mice significantly reduced kainate-induced zinc accumulation, neuronal death, and epileptic seizures in the hippocampal CA1 region [14]. Intracerebroventricular injection of CaEDTA reduced kainate-induced hippocampal neuronal death and epileptic seizures in wild-type mice, indicating that Zn2+ overload is a vital cause of neuronal death and epilepsy [14]. Knockout of Zip1 and Zip3 genes inhibited zinc absorption by pyramidal neurons in the hippocampus and kainic acid-induced epilepsy in mice, thus reducing neuronal damage [66]. Therefore, chelating Zn2+ may be a measure to prevent epileptic seizures. Transient receptor potential cation channel 5 (TRPC5) was involved in abnormal epileptic discharges and mediated neuronal death. Treatment of cortical neurons with hydrogen peroxide led to zinc-dependent TRPC5 opening, which triggered Ca2+ influx and caused neuronal death and seizures [67]. Therefore, zinc plays an important role in the process of seizure-mediated neuronal death. Ganaxolone (GX) inhibited seizures by acting on extrasynaptic GABA-A receptors. Zn2+ selectively blocked extrasynaptic GABA-A receptors in the hippocampus, thereby attenuating the antiepileptic effect of GX [68] (Figure 2).

5.5. Zinc Homeostasis Imbalance Promotes Alzheimer’s Disease Progression

Daily intraperitoneal injections of the Zn2+ chelator CQ for 2 weeks significantly decreased cognitive function in young mice [29]. Knockdown of ZnT3 significantly impaired cognitive function in mice [69]. Knockout of ZnT3 significantly reduced zinc levels in the mouse hippocampus, led to abnormal pre- and post-synaptic protein expressions, and affected learning and memory. CQ treatment reversed these phenomena in ZnT3 knockout mice, indicating that zinc homeostasis disorder led to cognitive impairment [70]. Intracellular and extracellular ROS increased the content of free Zn2+ [71]. ROS increased the content of Zn2+ in the neurons of AD mice. Intraperitoneal injection of pyruvate reduced the expression of ROS and the content of Zn2+ in neurons of AD mice and improved their cognitive function [72]. Therefore, pyruvate reduced the content of Zn2+ in the brain by downregulating the level of ROS in neurons, thereby improving the cognitive function of AD mice.
ZnAβ generated by the combination of Zn2+ and Aβ significantly inhibited long-term potentiation in the mouse hippocampus and activated microglia, and Zn2+ also enhanced the toxicity of Aβ to neuroblastoma cells [73]. Thus, ZnAβ is more neurotoxic than Aβ. Tau mice are transgenic mice containing human tau protein gene. High concentrations of zinc significantly increased the phosphorylation level of tau in tau mice, the number of tau tangles in the hippocampus, and memory impairment in tau mice [74]. Intracerebroventricular injection of ZnSO4 caused hyperphosphorylation of tau in the hippocampus, promoted oxidative stress injury, and impaired learning and memory function in rats. The nuclear factor E2 related factor 2 (Nrf2)/ heme oxygenase-1 (HO-1) pathway was inhibited by intracerebroventricular injection of ZnSO4 in rats and ZnSO4 treatment of SH-SY5Y cells [33]. In conclusion, ZnSO4 promoted tau phosphorylation by inhibiting the Nrf2/HO-1 pathway, thereby damaging the cognitive function of rats. Zn2+ enhanced tau neurotoxicity by inducing tau hyperphosphorylation and conformational changes. CQ reversed the effect of Zn2+ on tau, indicating that Zn2+ is the key to tau neurotoxicity [75]. Zn2+ enhanced the toxic effects of Aβ and promoted tau phosphorylation, indicating that restoring zinc homeostasis in the brain can prevent and treat AD [76] (Figure 2).

5.6. Zinc Led to Parkinson’s Disease Mediated by Inflammatory Response and Oxidative Stress

Intraperitoneal injection of high concentrations of ZnSO4 increased the content of Zn2+ in the substantia nigra and striatum, promoted oxidative stress, reduced the level of dopamine in the striatum, and reduced the spontaneous locomotor activity of rats [77]. Intraperitoneal injection of high concentrations of ZnSO4 in rats caused α-synuclein (α-syn) aggregation in the substantia nigra dopaminergic neurons [78]. Kufor-Rakeb syndrome (KRS) is a genetic disease caused by ATP13A2 (PARK9) mutations that manifests as Parkinson’s disease (PD). Knockdown of the PARK9 gene in primary cortical neurons caused the accumulation of α-syn, and Zn2+ treatment further increased α-syn accumulation, indicating that Zn2+ promoted the progression of PD [79]. Intraperitoneal injection of high concentrations of ZnSO4 in rats reduced the number of dopamine neurons and dopamine content, and increased the activity of NOX in the substantia nigra and striatum. The inhibition of microglial activation by minocycline reversed the effects of ZnSO4, suggesting that a high concentration of Zn2+ caused dopamine neuron damage by activating microglia to release inflammatory factors; thus, participating in the occurrence of PD [80]. The intraperitoneal injection of high concentrations of ZnSO4 in adult rats reduced the number of substantia nigra dopamine neurons and the content of striatal dopamine, increased the expression of oxidative stress and inflammatory factors, and caused PD symptoms. ZnSO4 treatment in newborn and adult rats further aggravated the above effects, indicating that a high concentration of Zn2+ promoted the occurrence and development of PD [81]. It was also found that high concentrations of Zn2+ promoted the release of proinflammatory factors from microglia in the substantia nigra by activating the NF-κB pathway and also promoted apoptosis [81] (Figure 2).
AMPA injection into the substantia nigra significantly increased the amount of free Zn2+ and the loss of dopamine neurons in the substantia nigra and caused movement disorder in rats. Zn2+ chelator reversed these effects, suggesting that AMPARCa-Zn mediated Zn2+ influx and damaged nigral dopamine neurons, leading to PD [82]. 6-OHDA impaired dopamine neurons in the substantia nigra and striatum by promoting extracellular Zn2+ to flow into dopamine neurons [83], while blocking Zn2+ influx into dopamine neurons inhibited paraquat-induced dopamine neuron loss [84]. In conclusion, a large influx of extracellular Zn2+ into dopaminergic neurons promoted the development of PD. Synaptic Zn2+ release led to a rapid increase in extracellular Zn2+ concentration, so maintaining synaptic zinc homeostasis is a potential strategy to combat PD [15,82].

5.7. Zinc Accumulation Promotes the Occurrence and Development of Multiple Sclerosis

ZnT3 on the synaptic vesicle membrane transports zinc into the vesicles for storage [20,21]. Knockdown of ZnT3 reduced Zn2+ release into the synaptic cleft, significantly inhibited spinal white matter demyelination, reduced the inflammatory response in mice with experimental autoimmune encephalomyelitis (EAE), and significantly inhibited the activation of MMP-9 and the formation of zinc plaques [3]. 1H10 is an inhibitor of AMP-activated protein kinase (AMPK) phosphorylation, and it was experimentally demonstrated that 1H10 also has the ability to chelate Zn2+. Intraperitoneal injection of 1H10 in mice inhibited EAE-induced demyelination and microglia/macrophage activation, inhibited EAE-induced MMP-9 activation and zinc plaque formation and alleviated multiple sclerosis (MS) [85]. Therefore, inhibition of AMPK phosphorylation or chelation of Zn2+ may be an effective target for MS treatment. Zn2+ damaged myelin sheath by activating MMP-9 to destroy the BBB and activating microglia to release proinflammatory cytokines [16] (Figure 2).

6. Summary and Outlook

Synaptic zinc homeostasis plays an important role in maintaining brain zinc homeostasis and CNS stability. When synaptic zinc is in a homeostatic state, preventive zinc supplementation can improve the body’s immunity and reduce the production of inflammatory factors in the brain and is conducive to the utilization of zinc in the body [12,13]. Under intense stress conditions, such as cerebral ischemia [8], epilepsy [14], and hypoglycemia [86], the presynaptic membrane Zn2+ is released at a great rate into the synaptic cleft, resulting in Zn2+ overload in the postsynaptic neurons and neurotoxicity. Excessive Zn2+ levels in the synaptic cleft also activate extracellular MMP and microglia, promoting neuroinflammation and causing tissue damage [16]. Thus, the use of extracellular Zn2+ chelators, reduces neuronal death caused by ischemic hypoxia and epilepsy, and resists neuroinflammation.
Neurodegenerative diseases are closely related to synaptic transmission, such as the imbalance of acetylcholine and dopamine levels in AD and PD patients. Zn2+ uptake in vitro and a large amount of Zn2+ released by impaired synaptic function can activate microglia and also enhance inflammatory protein (α-syn, tau, Aβ, and MMP) expression and neurotoxicity, thereby driving neuroinflammation-mediated neuronal damage [16,33,73,78]. Therefore, maintaining synaptic zinc homeostasis plays an important role in treating neurodegenerative diseases [11,15,16]. Synaptic zinc homeostasis disorders cause neurotoxicity and further affect nerve transmission, thereby promoting the progression of neurodegenerative diseases. Regulating zinc homeostasis in the early onset of neurodegenerative diseases may delay the disease progression. When neurodegenerative disease progresses to a moderate or severe degree, nerve transmission may become impaired, and zinc intake should be strictly controlled.
Preparations for zinc supplementation have evolved from inorganic zinc agents to organic zinc agents [87]. Inorganic zinc agents have the disadvantages of strong intestinal irritation and low bioavailability [88]. Now, an organic zinc agent is mainly used to improve zinc content in the body, including zinc gluconate, zinc citrate, zinc-enriched yeast and protein-chelated zinc [88,89,90]. The clinical application of zinc chelating agent is chloroiodohydroxyquine, which is mainly used in the treatment of dermatological diseases [91,92]. Some studies showed the neuroprotective effect of chelating Zn2+ after cerebral ischemia and hypoxia [8,31,63], and the development of a clinical drug to Zn2+ in the brain may have a high application prospect.
Neuroinflammation is a common pathophysiological basis of neurodegenerative diseases and is a major research topic in the field of neurodegenerative diseases [93]. Regulation of neuroinflammation is an important way to prevent and treat neurodegenerative diseases. Zinc homeostasis disorders affect the occurrence and development of neuroinflammation. Therefore, restoring zinc homeostasis by targeting specific proteins that regulate zinc homeostasis may be vital for the prevention and treatment of neuroinflammation-related neurodegenerative diseases. Hence, the role of zinc homeostasis disorders in the development and progression of neuroinflammation has received increasing attention. Further clarification of the molecular mechanism of neuroinflammation caused by zinc homeostasis disorders has profound scientific and clinical significance for the prevention and treatment of neuroinflammation-related neurodegenerative diseases.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 82060250), the Guangxi Science and Technology Plan Project (Grant No. AD20238035), the Natural Science Foundation of Hunan Province (Grant Nos. 2021JJ30614 and 2021JJ70036), and the Hunan Clinical Medical Technology Innovation and Guidance Project (Grant No. 2020SK517).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to all members of the research group for their discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Colonna, M.; Butovsky, O. Microglia function in the central nervous system during health and neurodegeneration. Annu. Rev. Immunol. 2017, 35, 441–468. [Google Scholar] [CrossRef] [PubMed]
  2. Orihuela, R.; McPherson, C.A.; Harry, G.J. Microglial M1/M2 polarization and metabolic states. Br. J. Pharmacol. 2016, 173, 649–665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Choi, B.Y.; Kim, I.Y.; Kim, J.H.; Kho, A.R.; Lee, S.H.; Lee, B.E.; Sohn, M.; Koh, J.Y.; Suh, S.W. Zinc transporter 3 (ZnT3) gene deletion reduces spinal cord white matter damage and motor deficits in a murine MOG-induced multiple sclerosis model. Neurobiol. Dis. 2016, 94, 205–212. [Google Scholar] [CrossRef] [PubMed]
  4. Feijó, G.D.S.; Jantsch, J.; Correia, L.L.; Eller, S.; Furtado-Filho, O.V.; Giovenardi, M.; Porawski, M.; Braganhol, E.; Guedes, R.P. Neuroinflammatory responses following zinc or branched-chain amino acids supplementation in obese rats. Metab. Brain Dis. 2022, 37, 1875–1886. [Google Scholar] [CrossRef]
  5. Lin, P.H.; Sermersheim, M.; Li, H.; Lee, P.H.U.; Steinberg, S.M.; Ma, J. Zinc in wound healing modulation. Nutrients 2017, 10, 16. [Google Scholar] [CrossRef] [Green Version]
  6. 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]
  7. Li, X.; Chen, S.; Mao, L.; Li, D.; Xu, C.; Tian, H.; Mei, X. Zinc Improves Functional recovery by regulating the secretion of granulocyte colony stimulating factor from microglia/macrophages after spinal cord injury. Front. Mol. Neurosci. 2019, 12, 18. [Google Scholar] [CrossRef] [PubMed]
  8. Ueba, Y.; Aratake, T.; Onodera, K.I.; Higashi, Y.; Hamada, T.; Shimizu, T.; Shimizu, S.; Yawata, T.; Nakamura, R.; Akizawa, T.; et al. Attenuation of zinc-enhanced inflammatory M1 phenotype of microglia by peridinin protects against short-term spatial-memory impairment following cerebral ischemia in mice. Biochem. Biophys. Res. Commun. 2018, 507, 476–483. [Google Scholar] [CrossRef]
  9. Doboszewska, U.; Młyniec, K.; Wlaź, A.; Poleszak, E.; Nowak, G.; Wlaź, P. Zinc signaling and epilepsy. Pharmacol. Ther. 2019, 193, 156–177. [Google Scholar] [CrossRef]
  10. da Silva, L.E.M.; de Santana, M.L.P.; Costa, P.R.F.; Pereira, E.M.; Nepomuceno, C.M.M.; Queiroz, V.A.O.; de Oliveira, L.P.M.; Machado, M.; de Sena, E.P. Zinc supplementation combined with antidepressant drugs for treatment of patients with depression: A systematic review and meta-analysis. Nutr. Rev. 2021, 79, 1–12. [Google Scholar] [CrossRef]
  11. Szewczyk, B. Zinc homeostasis and neurodegenerative disorders. Front. Aging Neurosci. 2013, 5, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. De Oliveira, S.; Feijó, G.D.S.; Neto, J.; Jantsch, J.; Braga, M.F.; Castro, L.; Giovenardi, M.; Porawski, M.; Guedes, R.P. Zinc supplementation decreases obesity-related neuroinflammation and improves metabolic function and memory in rats. Obesity 2021, 29, 116–124. [Google Scholar] [CrossRef] [PubMed]
  13. Mousaviyan, R.; Davoodian, N.; Alizadeh, F.; Ghasemi-Kasman, M.; Mousavi, S.A.; Shaerzadeh, F.; Kazemi, H. Zinc supplementation during pregnancy alleviates lipopolysaccharide-induced glial activation and inflammatory markers expression in a rat model of maternal immune activation. Biol. Trace Elem. Res. 2021, 199, 4193–4204. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, J.Y.; Kim, J.H.; Palmiter, R.D.; Koh, J.Y. Zinc released from metallothionein-iii may contribute to hippocampal CA1 and thalamic neuronal death following acute brain injury. Exp. Neurol. 2003, 184, 337–347. [Google Scholar] [CrossRef] [PubMed]
  15. Sikora, J.; Ouagazzal, A.M. Synaptic zinc: An emerging player in parkinson’s disease. Int. J. Mol. Sci. 2021, 22, 4724. [Google Scholar] [CrossRef]
  16. Choi, B.Y.; Jung, J.W.; Suh, S.W. The Emerging role of zinc in the pathogenesis of multiple sclerosis. Int J Mol Sci. 2017, 18, 2070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Kambe, T.; Tsuji, T.; Hashimoto, A.; Itsumura, N. The physiological, biochemical, and molecular roles of zinc transporters in zinc homeostasis and metabolism. Physiol. Rev. 2015, 95, 749–784. [Google Scholar] [CrossRef] [Green Version]
  18. Kimura, T.; Kambe, T. The functions of metallothionein and ZIP and ZnT transporters: An overview and perspective. Int. J. Mol. Sci. 2016, 17, 336. [Google Scholar] [CrossRef] [Green Version]
  19. Baltaci, A.K.; Yuce, K.; Mogulkoc, R. Zinc metabolism and metallothioneins. Biol. Trace Elem. Res. 2018, 183, 22–31. [Google Scholar] [CrossRef]
  20. Wenzel, H.J.; Cole, T.B.; Born, D.E.; Schwartzkroin, P.A.; Palmiter, R.D. Ultrastructural localization of zinc transporter-3 (ZnT-3) to synaptic vesicle membranes within mossy fiber boutons in the hippocampus of mouse and monkey. Proc. Natl. Acad. Sci. USA 1997, 94, 12676–12681. [Google Scholar] [CrossRef] [Green Version]
  21. Palmiter, R.D.; Cole, T.B.; Quaife, C.J.; Findley, S.D. ZnT-3, a putative transporter of zinc into synaptic vesicles. Proc. Natl. Acad. Sci. USA 1996, 93, 14934–14939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ohana, E.; Segal, D.; Palty, R.; Ton-That, D.; Moran, A.; Sensi, S.L.; Weiss, J.H.; Hershfinkel, M.; Sekler, I. A sodium zinc exchange mechanism is mediating extrusion of zinc in mammalian cells. J. Biol. Chem. 2004, 279, 4278–4284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Granzotto, A.; Sensi, S.L. Intracellular zinc is a critical intermediate in the excitotoxic cascade. Neurobiol. Dis. 2015, 81, 25–37. [Google Scholar] [CrossRef] [PubMed]
  24. Bitanihirwe, B.K.Y.; Woo, T.W. A conceptualized model linking matrix metalloproteinase-9 to schizophrenia pathogenesis. Schizophr. Res. 2020, 218, 28–35. [Google Scholar] [CrossRef]
  25. Ciccone, L.; Vandooren, J.; Nencetti, S.; Orlandini, E. Natural marine and terrestrial compounds as modulators of matrix metalloproteinases-2 (MMP-2) and MMP-9 in Alzheimer’s disease. Pharmaceuticals 2021, 14, 86. [Google Scholar] [CrossRef] [PubMed]
  26. Qin, W.; Li, J.; Zhu, R.; Gao, S.; Fan, J.; Xia, M.; Zhao, R.C.; Zhang, J. Melatonin protects blood-brain barrier integrity and permeability by inhibiting matrix metalloproteinase-9 via the NOTCH3/NF-κB pathway. Aging 2019, 11, 11391–11415. [Google Scholar] [CrossRef]
  27. Pilat, D.; Paumier, J.M.; García-González, L.; Louis, L.; Stephan, D.; Manrique, C.; Khrestchatisky, M.; Di Pasquale, E.; Baranger, K.; Rivera, S. MT5-MMP promotes neuroinflammation, neuronal excitability and Aβ production in primary neuron/astrocyte cultures from the 5xFAD mouse model of Alzheimer’s disease. J. Neuroinflamm. 2022, 19, 65. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, C.C.; Hsiao, L.D.; Tseng, H.C.; Kuo, C.M.; Yang, C.M. Pristimerin inhibits MMP-9 expression and cell migration through attenuating NOX/ROS-dependent NF-κB activation in rat brain astrocytes challenged with LPS. J. Inflamm. Res. 2020, 13, 325–341. [Google Scholar] [CrossRef] [PubMed]
  29. Frazzini, V.; Granzotto, A.; Bomba, M.; Massetti, N.; Castelli, V.; d’Aurora, M.; Punzi, M.; Iorio, M.; Mosca, A.; Delli Pizzi, S.; et al. The pharmacological perturbation of brain zinc impairs BDNF-related signaling and the cognitive performances of young mice. Sci. Rep. 2018, 8, 9768. [Google Scholar] [CrossRef] [Green Version]
  30. Ruan, Z.; Zhang, D.; Huang, R.; Sun, W.; Hou, L.; Zhao, J.; Wang, Q. Microglial activation damages dopaminergic neurons through MMP-2/-9-mediated increase of blood-brain barrier permeability in a Parkinson’s disease mouse model. Int. J. Mol. Sci. 2022, 23, 2793. [Google Scholar] [CrossRef]
  31. Higashi, Y.; Aratake, T.; Shimizu, S.; Shimizu, T.; Nakamura, K.; Tsuda, M.; Yawata, T.; Ueba, T.; Saito, M. Influence of extracellular zinc on M1 microglial activation. Sci. Rep. 2017, 7, 43778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Mortadza, S.S.; Sim, J.A.; Stacey, M.; Jiang, L.H. Signalling mechanisms mediating Zn(2+)-induced TRPM2 channel activation and cell death in microglial cells. Sci. Rep. 2017, 7, 45032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Lai, C.; Chen, Z.; Ding, Y.; Chen, Q.; Su, S.; Liu, H.; Ni, R.; Tang, Z. Rapamycin attenuated zinc-induced tau phosphorylation and oxidative stress in rats: Involvement of dual mTOR/p70S6K and Nrf2/HO-1 pathways. Front. Immunol. 2022, 13, 782434. [Google Scholar] [CrossRef] [PubMed]
  34. Moriyama, M.; Fujitsuka, S.; Kawabe, K.; Takano, K.; Nakamura, Y. Zinc potentiates lipopolysaccharide-induced nitric oxide production in cultured primary rat astrocytes. Neurochem. Res. 2018, 43, 363–374. [Google Scholar] [CrossRef] [PubMed]
  35. Higashi, Y.; Segawa, S.; Matsuo, T.; Nakamura, S.; Kikkawa, Y.; Nishida, K.; Nagasawa, K. Microglial zinc uptake via zinc transporters induces ATP release and the activation of microglia. Glia 2011, 59, 1933–1945. [Google Scholar] [CrossRef] [PubMed]
  36. Doi, T.; Hara, H.; Kajita, M.; Kamiya, T.; Adachi, T. Zinc regulates expression of IL-23 p19 mRNA via activation of eIF2α/ATF4 axis in HAPI cells. Biometals 2015, 28, 891–902. [Google Scholar] [CrossRef]
  37. Hongxia, L.; Yuxiao, T.; Zhilei, S.; Yan, S.; Yicui, Q.; Jiamin, S.; Xin, X.; Jianxin, Y.; Fengfeng, M.; Hui, S. Zinc inhibited LPS-induced inflammatory responses by upregulating A20 expression in microglia BV2 cells. J. Affect Disord. 2019, 249, 136–142. [Google Scholar] [CrossRef]
  38. Aratake, T.; Higashi, Y.; Ueba, Y.; Hamada, T.; Shimizu, T.; Shimizu, S.; Yawata, T.; Ueba, T.; Saito, M. The inhibitory role of intracellular free zinc in the regulation of Arg-1 expression in interleukin-4-induced activation of M2 microglia. Metallomics 2018, 10, 1501–1509. [Google Scholar] [CrossRef]
  39. Summersgill, H.; England, H.; Lopez-Castejon, G.; Lawrence, C.B.; Luheshi, N.M.; Pahle, J.; Mendes, P.; Brough, D. Zinc depletion regulates the processing and secretion of IL-1β. Cell Death Dis. 2014, 5, e1040. [Google Scholar] [CrossRef] [Green Version]
  40. Rivers-Auty, J.; Tapia, V.S.; White, C.S.; Daniels, M.J.D.; Drinkall, S.; Kennedy, P.T.; Spence, H.G.; Yu, S.; Green, J.P.; Hoyle, C.; et al. Zinc status alters Alzheimer’s disease progression through NLRP3-dependent inflammation. J. Neurosci. 2021, 41, 3025–3038. [Google Scholar] [CrossRef]
  41. Wong, C.P.; Magnusson, K.R.; Ho, E. Increased inflammatory response in aged mice is associated with age-related zinc deficiency and zinc transporter dysregulation. J. Nutr. Biochem. 2013, 24, 353–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sauer, A.K.; Grabrucker, A.M. Zinc deficiency during pregnancy leads to altered microbiome and elevated inflammatory markers in mice. Front. Neurosci. 2019, 13, 1295. [Google Scholar] [CrossRef] [PubMed]
  43. Sauer, A.K.; Malijauskaite, S.; Meleady, P.; Boeckers, T.M.; McGourty, K.; Grabrucker, A.M. Zinc is a key regulator of gastrointestinal development, microbiota composition and inflammation with relevance for autism spectrum disorders. Cell Mol. Life Sci. 2021, 79, 46. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, Z.; Qian, K.; Chen, J.; Qi, Y.; E, Y.; Liang, J.; Zhao, L. A biomimetic zeolite-based nanoenzyme contributes to neuroprotection in the neurovascular unit after ischaemic stroke via efficient removal of zinc and ROS. Acta Biomater. 2022, 144, 142–156. [Google Scholar] [CrossRef]
  45. Segawa, S.; Nishiura, T.; Furuta, T.; Ohsato, Y.; Tani, M.; Nishida, K.; Nagasawa, K. Zinc is released by cultured astrocytes as a gliotransmitter under hypoosmotic stress-loaded conditions and regulates microglial activity. Life Sci. 2014, 94, 137–144. [Google Scholar] [CrossRef]
  46. Olasehinde, T.A.; Olaniran, A.O.; Okoh, A.I. Neuroprotective effects of some seaweeds against Zn-induced neuronal damage in HT-22 cells via modulation of redox imbalance, inhibition of apoptosis and acetylcholinesterase activity. Metab. Brain Dis. 2019, 34, 1615–1627. [Google Scholar] [CrossRef]
  47. Hong, D.K.; Kho, A.R.; Lee, S.H.; Jeong, J.H.; Kang, B.S.; Kang, D.H.; Park, M.K.; Park, K.H.; Lim, M.S.; Choi, B.Y.; et al. Transient receptor potential melastatin 2 (TRPM2) inhibition by antioxidant, N-acetyl-l-cysteine, reduces global cerebral ischemia-induced neuronal death. Int. J. Mol. Sci. 2020, 21, 6026. [Google Scholar] [CrossRef]
  48. Güzel, M.; Nazıroğlu, M.; Akpınar, O.; Çınar, R. Interferon gamma-mediated oxidative stress induces apoptosis, neuroinflammation, zinc ion influx, and TRPM2 channel activation in neuronal cell line: Modulator role of curcumin. Inflammation 2021, 44, 1878–1894. [Google Scholar] [CrossRef]
  49. Zhao, X.; Sun, J.; Yuan, Y.; Lin, S.; Lin, J.; Mei, X. Zinc promotes microglial autophagy through NLRP3 inflammasome inactivation via XIST/miR-374a-5p axis in spinal cord injury. Neurochem. Res. 2022, 47, 372–381. [Google Scholar] [CrossRef]
  50. Pang, W.; Leng, X.; Lu, H.; Yang, H.; Song, N.; Tan, L.; Jiang, Y.; Guo, C. Depletion of intracellular zinc induces apoptosis of cultured hippocampal neurons through suppression of ERK signaling pathway and activation of caspase-3. Neurosci. Lett. 2013, 552, 140–145. [Google Scholar] [CrossRef] [PubMed]
  51. Li, D.; Wang, G.; Han, D.; Bi, J.; Li, C.; Wang, H.; Liu, Z.; Gao, W.; Gao, K.; Yao, T.; et al. MP resulting in autophagic cell death of microglia through zinc changes against spinal cord injury. Biomed Res. Int. 2016, 2016, 6090316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lin, J.Q.; Tian, H.; Zhao, X.G.; Lin, S.; Li, D.Y.; Liu, Y.Y.; Xu, C.; Mei, X.F. Zinc provides neuroprotection by regulating NLRP3 inflammasome through autophagy and ubiquitination in a spinal contusion injury model. CNS Neurosci. Ther. 2021, 27, 413–425. [Google Scholar] [CrossRef] [PubMed]
  53. Swardfager, W.; Herrmann, N.; Mazereeuw, G.; Goldberger, K.; Harimoto, T.; Lanctôt, K.L. Zinc in depression: A meta-analysis. Biol. Psychiatry 2013, 74, 872–878. [Google Scholar] [CrossRef] [PubMed]
  54. Siwek, M.; Dudek, D.; Paul, I.A.; Sowa-Kućma, M.; Zieba, A.; Popik, P.; Pilc, A.; Nowak, G. Zinc supplementation augments efficacy of imipramine in treatment resistant patients: A double blind, placebo-controlled study. J. Affect Disord. 2009, 118, 187–195. [Google Scholar] [CrossRef] [PubMed]
  55. Nowak, G.; Siwek, M.; Dudek, D.; Zieba, A.; Pilc, A. Effect of zinc supplementation on antidepressant therapy in unipolar depression: A preliminary placebo-controlled study. Pol. J. Pharmacol. 2003, 55, 1143–1147. [Google Scholar] [PubMed]
  56. Młyniec, K.; Davies, C.L.; Budziszewska, B.; Opoka, W.; Reczyński, W.; Sowa-Kućma, M.; Doboszewska, U.; Pilc, A.; Nowak, G. Time course of zinc deprivation-induced alterations of mice behavior in the forced swim test. Pharmacol. Rep. 2012, 64, 567–575. [Google Scholar] [CrossRef]
  57. Młyniec, K.; Nowak, G. Zinc deficiency induces behavioral alterations in the tail suspension test in mice. Effect of antidepressants. Pharmacol. Rep. 2012, 64, 249–255. [Google Scholar] [CrossRef]
  58. Tassabehji, N.M.; Corniola, R.S.; Alshingiti, A.; Levenson, C.W. Zinc deficiency induces depression-like symptoms in adult rats. Physiol. Behav. 2008, 95, 365–369. [Google Scholar] [CrossRef]
  59. Choi, B.Y.; Kim, I.Y.; Kim, J.H.; Lee, B.E.; Lee, S.H.; Kho, A.R.; Sohn, M.; Suh, S.W. Zinc plus cyclo-(His-Pro) promotes hippocampal neurogenesis in rats. Neuroscience 2016, 339, 634–643. [Google Scholar] [CrossRef]
  60. Yang, M.; Bao, D.; Shi, A.; Yuan, H.; Wang, J.; He, W.; Tong, X.; Qin, H. Zinc promotes patient-derived induced pluripotent stem cell neural differentiation via ERK-STAT signaling. Stem. Cells Dev. 2020, 29, 863–875. [Google Scholar] [CrossRef]
  61. Koh, J.Y.; Suh, S.W.; Gwag, B.J.; He, Y.Y.; Hsu, C.Y.; Choi, D.W. The role of zinc in selective neuronal death after transient global cerebral ischemia. Science 1996, 272, 1013–1016. [Google Scholar] [CrossRef] [PubMed]
  62. 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] [Green Version]
  63. Feng, L.; Gao, J.; Wang, Y.; Cheong, Y.K.; Ren, G.; Yang, Z. Etidronate-zinc complex ameliorated cognitive and synaptic plasticity impairments in 2-vessel occlusion model rats by reducing neuroinflammation. Neuroscience 2018, 390, 206–217. [Google Scholar] [CrossRef] [PubMed]
  64. Saghazadeh, A.; Mahmoudi, M.; Meysamie, A.; Gharedaghi, M.; Zamponi, G.W.; Rezaei, N. Possible role of trace elements in epilepsy and febrile seizures: A meta-analysis. Nutr. Rev. 2015, 73, 760–779. [Google Scholar] [CrossRef]
  65. Staley, K. Molecular mechanisms of epilepsy. Nat. Neurosci. 2015, 18, 367–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Qian, J.; Xu, K.; Yoo, J.; Chen, T.T.; Andrews, G.; Noebels, J.L. Knockout of Zn transporters Zip-1 and Zip-3 attenuates seizure-induced CA1 neurodegeneration. J. Neurosci. 2011, 31, 97–104. [Google Scholar] [CrossRef] [Green Version]
  67. Park, S.E.; Song, J.H.; Hong, C.; Kim, D.E.; Sul, J.W.; Kim, T.Y.; Seo, B.R.; So, I.; Kim, S.Y.; Bae, D.J.; et al. Contribution of zinc-dependent delayed calcium influx via TRPC5 in oxidative neuronal death and its prevention by novel TRPC antagonist. Mol. Neurobiol. 2019, 56, 2822–2835. [Google Scholar] [CrossRef] [Green Version]
  68. Chuang, S.H.; Reddy, D.S. Zinc reduces antiseizure activity of neurosteroids by selective blockade of extrasynaptic GABA-A receptor-mediated tonic inhibition in the hippocampus. Neuropharmacology 2019, 148, 244–256. [Google Scholar] [CrossRef] [PubMed]
  69. 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] [Green Version]
  70. 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] [Green Version]
  71. Aras, M.A.; Aizenman, E. Redox regulation of intracellular zinc: Molecular signaling in the life and death of neurons. Antioxid. Redox. Signal 2011, 15, 2249–2263. [Google Scholar] [CrossRef] [Green Version]
  72. Isopi, E.; Granzotto, A.; Corona, C.; Bomba, M.; Ciavardelli, D.; Curcio, M.; Canzoniero, L.M.; Navarra, R.; Lattanzio, R.; Piantelli, M.; et al. Pyruvate prevents the development of age-dependent cognitive deficits in a mouse model of Alzheimer’s disease without reducing amyloid and tau pathology. Neurobiol. Dis. 2015, 81, 214–224. [Google Scholar] [CrossRef] [PubMed]
  73. Lee, M.C.; Yu, W.C.; Shih, Y.H.; Chen, C.Y.; Guo, Z.H.; Huang, S.J.; Chan, J.C.C.; Chen, Y.R. Zinc ion rapidly induces toxic, off-pathway amyloid-β oligomers distinct from amyloid-β derived diffusible ligands in Alzheimer’s disease. Sci. Rep. 2018, 8, 4772. [Google Scholar] [CrossRef]
  74. Craven, K.M.; Kochen, W.R.; Hernandez, C.M.; Flinn, J.M. Zinc exacerbates tau pathology in a tau mouse model. J. Alzheimers Dis. 2018, 64, 617–630. [Google Scholar] [CrossRef]
  75. Huang, Y.; Wu, Z.; Cao, Y.; Lang, M.; Lu, B.; Zhou, B. Zinc binding directly regulates tau toxicity independent of tau hyperphosphorylation. Cell Rep. 2014, 8, 831–842. [Google Scholar] [CrossRef] [Green Version]
  76. Huat, T.J.; Camats-Perna, J.; Newcombe, E.A.; Valmas, N.; Kitazawa, M.; Medeiros, R. Metal toxicity links to Alzheimer’s disease and neuroinflammation. J. Mol. Biol. 2019, 431, 1843–1868. [Google Scholar] [CrossRef] [PubMed]
  77. Kumar, A.; Ahmad, I.; Shukla, S.; Singh, B.K.; Patel, D.K.; Pandey, H.P.; Singh, C. Effect of zinc and paraquat co-exposure on neurodegeneration: Modulation of oxidative stress and expression of metallothioneins, toxicant responsive and transporter genes in rats. Free Radic Res. 2010, 44, 950–965. [Google Scholar] [CrossRef]
  78. Kumar, V.; Singh, D.; Singh, B.K.; Singh, S.; Mittra, N.; Jha, R.R.; Patel, D.K.; Singh, C. Alpha-synuclein aggregation, Ubiquitin proteasome system impairment, and L-Dopa response in zinc-induced Parkinsonism: Resemblance to sporadic Parkinson’s disease. Mol. Cell Biochem. 2018, 444, 149–160. [Google Scholar] [CrossRef]
  79. Tsunemi, T.; Krainc, D. Zn+2 dyshomeostasis caused by loss of ATP13A2/PARK9 leads to lysosomal dysfunction and alpha-synuclein accumulation. Hum. Mol. Genet. 2014, 23, 2791–2801. [Google Scholar] [CrossRef] [Green Version]
  80. Kumar, V.; Singh, B.K.; Chauhan, A.K.; Singh, D.; Patel, D.K.; Singh, C. Minocycline rescues from zinc-induced nigrostriatal dopaminergic neurodegeneration: Biochemical and molecular interventions. Mol. Neurobiol. 2016, 53, 2761–2777. [Google Scholar] [CrossRef] [PubMed]
  81. Mittra, N.; Chauhan, A.K.; Singh, G.; Patel, D.K.; Singh, C. Postnatal zinc or paraquat administration increases paraquat or zinc-induced loss of dopaminergic neurons: Insight into augmented neurodegeneration. Mol. Cell Biochem. 2020, 467, 27–43. [Google Scholar] [CrossRef] [PubMed]
  82. Tamano, H.; Morioka, H.; Nishio, R.; Takeuchi, A.; Takeda, A. AMPA-induced extracellular Zn(2+) influx into nigral dopaminergic neurons causes movement disorder in rats. Neurotoxicology 2018, 69, 23–28. [Google Scholar] [CrossRef]
  83. Tamano, H.; Nishio, R.; Morioka, H.; Takeda, A. Extracellular Zn(2+) influx into nigral dopaminergic neurons plays a key role for pathogenesis of 6-Hydroxydopamine-induced Parkinson’s disease in rats. Mol. Neurobiol. 2019, 56, 435–443. [Google Scholar] [CrossRef] [PubMed]
  84. Tamano, H.; Morioka, H.; Nishio, R.; Takeuchi, A.; Takeda, A. Blockade of rapid influx of extracellular Zn(2+) into nigral dopaminergic neurons overcomes paraquat-induced Parkinson’s disease in rats. Mol. Neurobiol. 2019, 56, 4539–4548. [Google Scholar] [CrossRef] [PubMed]
  85. Choi, B.Y.; Jeong, J.H.; Eom, J.W.; Koh, J.Y.; Kim, Y.H.; Suh, S.W. A novel zinc chelator, 1H10, ameliorates experimental autoimmune encephalomyelitis by modulating zinc toxicity and AMPK activation. Int. J. Mol. Sci. 2020, 21, 3375. [Google Scholar] [CrossRef] [PubMed]
  86. Kho, A.R.; Choi, B.Y.; Kim, J.H.; Lee, S.H.; Hong, D.K.; Lee, S.H.; Jeong, J.H.; Sohn, M.; Suh, S.W. Prevention of hypoglycemia-induced hippocampal neuronal death by N-acetyl-L-cysteine (NAC). Amino Acids 2017, 49, 367–378. [Google Scholar] [CrossRef] [PubMed]
  87. Zhang, L.; Guo, Q.; Duan, Y.; Lin, X.; Ni, H.; Zhou, C.; Li, F. Comparison of the Effects of Inorganic or Amino Acid-Chelated Zinc on Mouse Myoblast Growth in vitro and Growth Performance and Carcass Traits in Growing-Finishing Pigs. Front. Nutr. 2022, 9, 857393. [Google Scholar] [CrossRef]
  88. Fazel Torshizi, F.; Chamani, M.; Khodaei, H.R.; Sadeghi, A.A.; Hejazi, S.H.; Majidzadeh Heravi, R. Therapeutic effects of organic zinc on reproductive hormones, insulin resistance and mTOR expression, as a novel component, in a rat model of Polycystic ovary syndrome. Iran J. Basic Med. Sci. 2020, 23, 36–45. [Google Scholar] [CrossRef]
  89. Maares, M.; Keil, C.; Pallasdies, L.; Schmacht, M.; Senz, M.; Nissen, J.; Kieserling, H.; Drusch, S.; Haase, H. Zinc availability from zinc-enriched yeast studied with an in vitro digestion/Caco-2 cell culture model. J. Trace Elem. Med. Biol. 2022, 71, 126934. [Google Scholar] [CrossRef] [PubMed]
  90. Zhang, Z.; Cheng, Q.; Liu, Y.; Peng, C.; Wang, Z.; Ma, H.; Liu, D.; Wang, L.; Wang, C. Zinc-Enriched Yeast May Improve Spermatogenesis by Regulating Steroid Production and Antioxidant Levels in Mice. Biol. Trace Elem. Res. 2022, 200, 3712–3722. [Google Scholar] [CrossRef]
  91. Da Costa, B.; Pippi, B.; Berlitz, S.J.; Carvalho, A.R.; Teixeira, M.L.; Külkamp-Guerreiro, I.C.; Andrade, S.F.; Fuentefria, A.M. Evaluation of activity and toxicity of combining clioquinol with ciclopirox and terbinafine in alternative models of dermatophytosis. Mycoses 2021, 64, 727–733. [Google Scholar] [CrossRef] [PubMed]
  92. Da Costa, B.; Pippi, B.; Andrzejewski Kaminski, T.F.; Andrade, S.F.; Fuentefria, A.M. In vitro antidermatophytic synergism of double and triple combination of clioquinol with ciclopirox and terbinafine. Mycoses 2020, 63, 993–1001. [Google Scholar] [CrossRef] [PubMed]
  93. Kwon, H.S.; Koh, S.H. Neuroinflammation in neurodegenerative disorders: The roles of microglia and astrocytes. Transl. Neurodegener 2020, 9, 42. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Zinc homeostasis disorder in the brain promotes neuroinflammation. When diseases, such as AD and PD, and conditions, such as cerebral ischemia and hypoxia, lead to synaptic function disorder, synaptic zinc levels are in a state of fragile balance or imbalance. In this case, zinc supplementation increases the extracellular Zn2+ and synaptic Zn2+ content in the brain, leading to the uncontrolled release of a large amount of Zn2+ from the synapse to the extracellular space. Large amounts of extracellular Zn2+ damage neurons in several ways: 1. Zn2+ flows directly into neurons, thus promoting oxidative stress and causing neuronal death; 2. Zn2+ damages neurons by activating microglia to produce proinflammatory factors and ROS; 3. Zn2+ damages neurons by enhancing the expression and neurotoxicity of inflammatory proteins (α-syn, Tau, Aβ, and MMP). Therefore, patients with disorders involving abnormal synaptic function should strictly control zinc uptake. α-synuclein (α-syn); Alzheimer’s disease (AD); Parkinson’s disease (PD); β-amyloid (Aβ); matrix metalloproteinase (MMP); interleukin (IL); tumor necrosis factor-α (TNF-α); reactive oxygen species (ROS); zinc transporter (ZnT); Zrt/Irt-like protein (ZIP); Ca2+-and Zn2+-permeable GluR2-lacking AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid) receptors (AMPARCa-Zn); N-methyl-d-aspartate receptor (NMDAR); voltage-gated calcium channel (VGCC); metallothionein (MT).
Figure 1. Zinc homeostasis disorder in the brain promotes neuroinflammation. When diseases, such as AD and PD, and conditions, such as cerebral ischemia and hypoxia, lead to synaptic function disorder, synaptic zinc levels are in a state of fragile balance or imbalance. In this case, zinc supplementation increases the extracellular Zn2+ and synaptic Zn2+ content in the brain, leading to the uncontrolled release of a large amount of Zn2+ from the synapse to the extracellular space. Large amounts of extracellular Zn2+ damage neurons in several ways: 1. Zn2+ flows directly into neurons, thus promoting oxidative stress and causing neuronal death; 2. Zn2+ damages neurons by activating microglia to produce proinflammatory factors and ROS; 3. Zn2+ damages neurons by enhancing the expression and neurotoxicity of inflammatory proteins (α-syn, Tau, Aβ, and MMP). Therefore, patients with disorders involving abnormal synaptic function should strictly control zinc uptake. α-synuclein (α-syn); Alzheimer’s disease (AD); Parkinson’s disease (PD); β-amyloid (Aβ); matrix metalloproteinase (MMP); interleukin (IL); tumor necrosis factor-α (TNF-α); reactive oxygen species (ROS); zinc transporter (ZnT); Zrt/Irt-like protein (ZIP); Ca2+-and Zn2+-permeable GluR2-lacking AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid) receptors (AMPARCa-Zn); N-methyl-d-aspartate receptor (NMDAR); voltage-gated calcium channel (VGCC); metallothionein (MT).
Biomolecules 13 00416 g001
Figure 2. Relationship between synaptic zinc and neuroinflammation. Zinc is an indispensable trace element for maintaining life-preserving activities of organisms. When synaptic zinc is in homeostasis, the appropriate zinc supplementation enhances the body’s utilization of zinc, thereby enhancing the body’s ability to resist neuroinflammation. However, synaptic zinc disorder further damages the synaptic function when suffering from synapse damage-related diseases. Zn2+ uptake in vitro and the large amount of Zn2+ released by impaired synapses promote neuroinflammation, so maintaining synaptic zinc homeostasis inhibits neuroinflammation. ERK (extracellular signal-regulated kinase); granulocyte colony stimulating factor (G-CSF); nuclear factor-κB (NF-κB); NOD-like receptor protein 3 (NLRP3); interleukin (IL); tumor necrosis factor-α (TNF-α); prefrontal cortex (PFC); lipopolysaccharide (LPS); Alzheimer’s disease (AD); Parkinson’s disease (PD); multiple sclerosis (MS); cerebral ischemia and hypoxia (CIH); epilepsy (EP); β-amyloid (Aβ); α-synuclein (α-syn); reduced nicotinamide adenine dinucleotide phosphate oxidase (NOX); matrix metalloproteinase (MMP); reactive oxygen species (ROS).
Figure 2. Relationship between synaptic zinc and neuroinflammation. Zinc is an indispensable trace element for maintaining life-preserving activities of organisms. When synaptic zinc is in homeostasis, the appropriate zinc supplementation enhances the body’s utilization of zinc, thereby enhancing the body’s ability to resist neuroinflammation. However, synaptic zinc disorder further damages the synaptic function when suffering from synapse damage-related diseases. Zn2+ uptake in vitro and the large amount of Zn2+ released by impaired synapses promote neuroinflammation, so maintaining synaptic zinc homeostasis inhibits neuroinflammation. ERK (extracellular signal-regulated kinase); granulocyte colony stimulating factor (G-CSF); nuclear factor-κB (NF-κB); NOD-like receptor protein 3 (NLRP3); interleukin (IL); tumor necrosis factor-α (TNF-α); prefrontal cortex (PFC); lipopolysaccharide (LPS); Alzheimer’s disease (AD); Parkinson’s disease (PD); multiple sclerosis (MS); cerebral ischemia and hypoxia (CIH); epilepsy (EP); β-amyloid (Aβ); α-synuclein (α-syn); reduced nicotinamide adenine dinucleotide phosphate oxidase (NOX); matrix metalloproteinase (MMP); reactive oxygen species (ROS).
Biomolecules 13 00416 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, S.; Wang, N.; Long, Y.; Wu, Z.; Zhou, S. Zinc Homeostasis: An Emerging Therapeutic Target for Neuroinflammation Related Diseases. Biomolecules 2023, 13, 416. https://doi.org/10.3390/biom13030416

AMA Style

Liu S, Wang N, Long Y, Wu Z, Zhou S. Zinc Homeostasis: An Emerging Therapeutic Target for Neuroinflammation Related Diseases. Biomolecules. 2023; 13(3):416. https://doi.org/10.3390/biom13030416

Chicago/Turabian Style

Liu, Shunfeng, Nan Wang, Yaqian Long, Zhuan Wu, and Shouhong Zhou. 2023. "Zinc Homeostasis: An Emerging Therapeutic Target for Neuroinflammation Related Diseases" Biomolecules 13, no. 3: 416. https://doi.org/10.3390/biom13030416

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