Skip to main content

Zinc transporters in Alzheimer’s disease

Abstract

Alzheimer’s disease (AD) is the most devastating neurodegenerative disorder. Due to the increase in population and longevity, incidence will triple by the middle of the twenty-first century. So far, no treatment has prevented or reversed the disease. More than 20 years of multidisciplinary studies have shown that brain zinc dyshomeostasis may play a critical role in AD progression, which provides encouraging clues for metal-targeted therapies in the treatment of AD. Unfortunately, the pilot clinical application of zinc chelator and/or ionophore strategy, such as the use of quinoline-based compounds, namely clioquinol and PBT2, has not yet been successful. The emerging findings revealed a list of key zinc transporters whose mRNA or protein levels were abnormally altered at different stages of AD brains. Furthermore, specifically modulating the expression of some of the zinc transporters in the central nervous system through genetic methods slowed down or prevented AD progression in animal models, resulting in significantly improved cognitive performance, movement, and prolonged lifespan. Although the underlying molecular mechanisms are not yet fully understood, it shed new light on the treatment or prevention of the disease. This review considers recent advances regarding AD, zinc and zinc transporters, recapitulating their relationships in extending our current understanding of the disease amelioration effects of zinc transport proteins as potential therapeutic targets to cure AD, and it may also provide new insights to identify novel therapeutic strategies for ageing and other neurodegenerative diseases, such as Huntington’s and Parkinson’s disease.

Introduction

Alzheimer’s disease (AD) is a progressive neurodegenerative disease, which leaves many elderly people cognitively impaired [1] and the numbers are expected to reach 131 million by 2050 [2]. The disorder is characterized by a progressive loss of memory due to the degeneration and/or loss of neurons related to cognition in the hippocampus and cortex. Pathologically, the extracellular accumulation of amyloid plaques rich in the β-amyloid (Aβ) peptide and intracellular deposition of neurofibrillary tangles (NFTs) with hyperphosphorylated tau in the brain are the two major hallmarks of AD [3, 4]. Aβ aggregates can form complexes with redox active metals, and the phosphorylation of tau is also associated with metals [5], especially zinc, which interplays with the Aβ protein, aggravating the precipitation of senile plaques, and it can also bind to tau proteins, contributing to the formation of NFTs [6,7,8]. Considerable early clinical studies, including pharmaceutical approaches designed to change the level of zinc or to alter the distribution of zinc and copper, have provided encouraging clues for metal-targeted therapies in the treatment and prevention of AD [9,10,11].

Zinc metabolism is firmly controlled by zinc transporters that transport zinc through the membrane to the cytoplasm or organelles and maintain zinc homeostasis both at the cellular and systemic levels. Zinc transporters are classified into two families: the ZnT (SLC30) family and the ZIP (SLC39) family. To the best of our knowledge, most zinc transporters are also expressed in mammalian brains, such as ZnT1, ZnT3, ZnT10, ZIP1 and ZIP6 [12,13,14,15,16,17]. Moreover, it has been established that a considerable number of zinc transporters were abnormally expressed in the brains of AD patients and APP/PS1 mice [18,19,20], suggesting that zinc transporters may play a critical role in the progression of AD. Since so far there has been no effective therapy to halt or reverse AD, we propose that altering zinc homeostasis by modulating the expression or capability of zinc transporters may be a better way to ameliorate AD. This review focuses first on the relationship between zinc transporters and AD. After that, we summarize the function of zinc transporters that are involved in brain zinc homeostasis and assess recent evidence for treating AD with the alteration of zinc transporters.

The important role of zinc in the brain

As an indispensable trace element, zinc is a structural or functional component of many proteins. It also participates in various physiological functions, including but not limited to DNA synthesis, brain development and neurotransmission [21,22,23]. Considering its multifaceted role in biological processes, alterations in zinc concentrations from normal levels can induce many devastating diseases, e.g., zinc deficiency leading to mental lethargy, growth retardation and immune dysfunction, while zinc overload can affect the function of lymphocytes and cause copper deficiency [24].

Besides the important role of zinc in normal mammalian brain development and physiology [25], it also plays a critical role in the function of the central nervous system (CNS). There are three pools of zinc in the brain, namely vesicular zinc, protein-bound zinc and free zinc [26]. Vesicular zinc exists mainly in glutamatergic neurons, but not all glutamatergic neurons contain zinc [27, 28]. Studies found that zinc can permeate through the neuronal membrane via several ion channels, such as Ca2+-permeable α-amino-3-hydroxyl-5-methyl-4-isoxazolepropionate-acid (AMPA) /kainite (Ca2+-A/K) channels [29], high-voltage activated (HVA) Ca2+ channels [30], and N-methyl-D-aspartate (NMDA) receptor-gated channels [31], depending on neuronal stimulation. Zinc binds to the extracellular amino-terminal domain (ATD) of NMDA receptors [32], and activation of AMPA and NMDA receptors facilitated Zn2+ influx into neurons [31, 33]. Zinc can also enter neurons through voltage-gated calcium channels, especially when the pH is lower [30]. Toxic activation studies have shown that Ca2+-A/K channels can carry a greater flux of zinc than influx of zinc in comparison with voltage-sensitive Ca2+ channel (VSCCs) and NMDA channels, and the permeability of Ca2+-A/K channels to zinc is also comparable to their permeability for Ca2+ [29]. Because they are concentrated in postsynaptic regions of dendrites adjacent to sites of presynaptic release, the Ca-A/K channels may constitute the principal physiologic route for Zn2+ translocation [29]. Zinc functions as a regulator of many enzymes and can directly interact with mitochondria and cause reactive oxygen species (ROS) production leading to important physio-pathological consequences, including neurodegeneration [29, 34]. Moreover, there is increasing evidence that high level of intracellular free zinc can induce neuronal death by inhibiting the process of energy production through interacting with targets essential to mitochondrial electron transport, the tricarboxylic acid cycle (TCA), and glycolysis [35], and the consequences of zinc overload also include reduced cellular ATP levels and loss of mitochondrial membrane potential [35]. A later study through direct biochemical assays demonstrated that excess zinc can irreversibly damage the major enzymes of energy production and antioxidant defense prior to mitochondrial permeability transition [36]. Zinc will participate in numerous signal transduction processes as soon as it enters the postsynaptic neuron, and as a neuromodulator, zinc can influence the brain excitability by inhibiting excitatory and inhibitory receptors, thus Zn2+ qualifying as both a neurotransmitter and a second messenger [37] .

On the other hand, zinc can modulate synaptic plasticity in the form of long-term potentiation (LTP) and long-term depression (LTD), respectively, which are considered to be underlying learning and memory, and are differentially modulated by NMDA receptor subunits [37,38,39]. When stimulated, zinc is co-released with glutamate into the synaptic cleft to modulate the activity of post-synaptic proteins, NMDA and AMPA receptors. Simultaneously, zinc can transactivate the tropomyosin-related kinase B (TrkB) receptor in cultured cortical neurons. As soon as zinc enters the postsynaptic neuron, it activates brain-derived neurotrophic factor (BDNF) signaling in a neurotrophin-independent manner [40, 41]. Zinc can also affect BDNF signaling by promoting maturation of pro-BDNF to BDNF all through the activation of matrix metalloproteinases (MMPs), thereby significantly increasing BDNF level in brains of AD mice and greatly delaying hippocampal-dependent memory deficits [42]. A subsequent study has shown that under physiological conditions, it is BDNF rather than vesicular zinc that activates TrkB in hippocampal mossy fiber axons [43]. Synaptic release of Zn2+ activates a Zn2+-sensing receptor, mZnR/GPR39, induces Ca2+-signaling, then activates ERK1/2 MAPK and up-regulates clusterin; however, Zn2+ signaling via mZnR/GPR39 is disrupted by amyloid-β in AD brains, which is a critical pathological component of AD [44]. Such a vital role of zinc in the brain has prompted the scientific community to continue exploring the etiology and therapeutic strategies based on zinc dyshomeostasis in neurodegenerative diseases, especially for AD, Parkinson’s disease (PD) and Huntington’s disease (HD) [1, 37, 45].

The maintenance of zinc homeostasis in human and mammalian cells

Endogenous zinc equilibrium is promoted by zinc-binding proteins (metallothioneins, MTs) and zinc transporters. There are four MTs in human and mammalian cells: MT-1, MT-2, MT-3 and MT-4, which can bind up to seven divalent metals, including Zn2+, Cu2+, Mn2+, Fe2+ and Cd2+, etc. MT-1 and MT-2 are widely expressed, MT-3 is abundant in the brain, and MT-4 is expressed in some epithelial cells. When zinc is in excess, MTs may function in sequestration of excess zinc, thereby protecting cells from zinc toxicity [46]. Meanwhile, zinc transporters are responsible for the movement of zinc across the membrane; they either pump zinc out of the cytoplasm or load zinc into the intracellular space. Generally, zinc transporters are divided into two major families: the SLC30 family (ZnTs1–10) and the SLC39 family (ZIPs1–14) (see Fig. 1). The SLC30 family of zinc transporters liberates zinc out of the cytoplasm or into cellular organelles, resulting in the reduction of zinc concentration in the cytoplasm. In contrast, the SLC39 family of zinc transporters increase intracellular zinc level by moving zinc into the cytoplasm, either from the organelles or the extracellular space [14, 47].

Fig. 1
figure 1

An overview of the locations of some crucial zinc transporters in mammalian cells and the intracellular and intercellular compartmentation of zinc ions. Zinc transporters are classified into two major families, the ZnT (SLC30) family and the ZIP (SLC39) family. As indicated by arrows in the figure, the general functions of ZnTs (ZnT1–7) is to reduce cytoplasmic zinc levels through mobilization of zinc out of cells or into intracellular compartments from the cytoplasm, whereas the function of ZiPs (ZiP1, ZiP4, ZiP6, ZiP8–10, ZiP13–14) is to increase cytoplasmic zinc levels through absorbing zinc into cytoplasm from extracellular space or mobilizing zinc out of intracellular compartments

Most ZnT proteins have six transmembrane domains (TMDs), with a long His-rich loop between TMDs Ις and ς, which may function as a sensor of cytosolic zinc levels. The C- and N- termini of ZnTs are located inside the cytoplasm [14, 47]. ZnT proteins usually form homodimers, except ZnT5 and ZnT6, which form heterodimers to transport zinc to the secretory pathway, where zinc is loaded into the zinc-requiring enzymes [48].

ZIP proteins are distinct from ZnTs in many aspects. Most ZIP proteins have eight TMDs with a His-rich domain between TMDs ΙΙΙ and Ις, and hydrophilic residues in TMD ς that is thought to function as a metal sensor. The C-and N-termini of ZIP proteins faces the extracytoplasmic space [14, 47].

There is a growing body of evidence showing that ZnTs and ZIPs play a fundamental role in maintaining zinc homeostasis, which profoundly affects human health, either directly or indirectly. Mutations in ZnTs and ZIPs have been found to be closely associated with a considerable number of human diseases, such as ZIP4 in Acrodermatitis enteropathica (AE), ZIP13 in the spondylocheiro dysplastic form of Ehlers–Danlos syndrome (SCD-EDS), and ZnT10 in Parkinson’s Disease [47, 49]. Consequently, studies on ZnTs and ZIPs are receiving an increasing clinical interest.

Zinc transporters in the brain

The mechanism of zinc homeostasis in the brain is complex. Elevated or reduced serum zinc has little effect on the zinc content in the brain. This indicates that zinc homeostasis in the brain is rigorously controlled and is not easily altered by peripheral zinc. The primary group of proteins involved in maintaining zinc homeostasis in an elaborate manner are zinc transporters, and most of them demonstrate an age-dependent alteration in the brain [50,51,52]. Table 1 lists the expression of zinc transporters detected in human or mouse brains and their alterations in the progression of AD, including at least seven ZnT (ZnT1, ZnT3, ZnT4, ZnT5, ZnT6, ZnT7 and ZnT10) and six ZIP (ZIP1, ZIP3, ZIP4, ZIP6, ZIP9 and ZIP13) transporters.

Table 1 Summary of brain zinc transporters and their alterations in the progression of AD

ZnT subfamily members in the brain

ZnT1 was the first member identified by isolation from a rat kidney cDNA expression library [14]. It localizes to the plasma membrane and is ubiquitously expressed [66]. In the brains of mice, ZnT1 is highly expressed in dendritic spines and synaptic membranes [12]. Besides, ZnT1 interacts with GluN2A-containing NMDA receptors, forming GluN2A/ZnT1 complexes [67]. Thus, in addition to exporting zinc to extracellular space, ZnT1 plays an important role in the modulation of postsynaptic receptors.

ZnT3 and ZnT4 are both expressed in the brain, but in different regions. Acting as a regulator of cognition and learning, ZnT3 was cloned by screening of a mouse λ library based on its homology to the ZnT2 transporter [14]. ZnT3 localizes to the synaptic vesicular membrane and is abundant in the hippocampus and cortex. In the brains of ZnT3 knockout mice, zinc is nearly undetectable in synaptic vesicles, suggesting that ZnT3 is responsible for transporting zinc into vesicles [54, 68]. Furthermore, ZnT3 also exists in the choroid plexus epithelial cells, the Bergman glial cells of mouse cerebellar cortex and the postganglionic neurons of mouse superior cervical ganglion [13]. However, ZnT3 may exert some undiscovered functions in the brain. ZnT4 was initially identified as the vital gene involved in the lethal milk mutant [69], and was later detected abundantly in mammary brain and glands, especially in the lysosomal and endosomal compartments, but with low expression in other tissues [53]. Perhaps, ZnT4 mainly facilitates the entry of zinc into secretory vesicles.

ZnT5 has homology with ZRC1 in yeast, and is widely expressed in all human tissues. Interestingly, ZnT5 and ZnT6 are both restricted to the membrane of the Golgi apparatus in mammalian cells, including neurons. ZnT5 and ZnT6 form hetero-oligomeric complexes for transporting zinc to tissue-nonspecific alkaline phosphatase (TNAP) [48]. Another ZnT transporter, ZnT7, also exists in the membrane of the Golgi apparatus. ZnT7 has homology with ZnT1, which is ubiquitously expressed in mouse tissues, particularly in the liver and small intestine. It is also found in the brain, kidneys and lungs [58]. The precise roles of ZnT5, ZnT6 and ZnT7 are obscure. Perhaps ZnT7 interacts with ZnT5 and ZnT6 to transport zinc to the secretory pathway and controls zinc homeostasis in the neuron.

Currently, there is limited knowledge regarding ZnT10. ZnT10 is highly homologous to ZnT1 and is mainly expressed in the fetal brain and fetal liver tissues. Hence, it can be speculated that ZnT10 is a crucial factor in fetal development [14]. ZnT10 is associated with PD and AD [49, 60], but whether ZnT10 transports zinc or manganese is controversial. Under normal conditions of cultured neuroblastoma cell lines, ZnT10 localizes at the Golgi apparatus. However, if the extracellular zinc concentration is higher than 100 μM, ZnT10 translocates to the plasma membrane and its mRNA expression level decreases [70]. Interestingly, other researchers have found that ZnT10 may primarily transport manganese rather than zinc [49, 71, 72]. Structurally, ZnT10 lacks a His-rich region that is similar to the manganese transporter shCDF8 in Stylosanthes hamate [73], whereas mutations in ZnT10 cause manganese disturbance [49].

ZIP subfamily members in the brain

There are fewer studies on ZIPs relative to ZnTs. Human ZIP1 was identified through its homology with the Arabidopsis ZIP1 transporter, and is ubiquitously expressed in human tissues [74, 75]. Belloni-Olivi et al. [15] found ZIP1 to be abundant in some areas of the brain in rats, including the hippocampus and the thalamus. It was found that ZIP1 localizes to the membrane of microglia, which is responsible for zinc uptake in astrocyte and microglial cells [16]. Microglia are resident immune cells in the brain. Researchers have shown that zinc is a key factor in transient global ischemia-induced microglial activation through sequential activation of NADPH oxidase and Poly (ADP-ribose) polymerase-1 (PARP-1). Interestingly, ZIP1 is involved in this process [76].

It was also found that ZIP3 and ZIP4 are expressed in the brain. ZIP3 is highly expressed in the testes and mammary cells. Qian et al. [76] discovered that ZIP3 was abundant in hippocampus pyramidal neurons, and knockout of ZIP3 can attenuate seizure-induced CA1 degeneration. ZIP4 is widely expressed in human tissues, such as the small intestine, stomach and kidneys, and it plays a fundamental role in the absorption of zinc in the small intestine. In rats, researchers found ZIP4 to be expressed in the choroid plexus [15]. ZIP4 is considered a marker of glioma, due to its higher expression in higher grade of gliomas with shorter overall survival [63].

Functioning as a zinc transporter, ZIP6 is widely expressed, and is localized to the plasma membrane. In particular ZIP6 is found in the hippocampal pyramidal and the apical membrane of the choroid plexus. The researchers found that the mRNA level of ZIP6 increased in the brain of neonatal rats when zinc was deficient [17].

There are a few studies describing ZIP9 and ZIP13 in the brain. ZIP9 was found to regulate zinc homeostasis in the secretory pathway [64], and was characterized as a membrane androgen receptor (mAR) [77]. ZIP13 plays a key role in the development of bone, teeth and connective tissue, and its defect is responsible for spondylocheirodysplastic Ehlers-Danlos syndrome (SCD-EDS) [65]. Jeong et al. [42] reported that SCD-EDS was caused by zinc deficiency in the endoplasmic reticulum (ER) rather than an overload, because the evidence showed that ZIP13 functioned to release labile zinc from vesicular stores for use in the ER and other compartments. However, Xiao et al. [71] found that ZIP13 mediated iron export in Drosophila, suggesting that SCD-EDS may also be caused by the failure of the iron delivery to the secretory compartments [78]. A very recent study found that the expression of ZIP9 and ZIP13 increased significantly with aging in the normal human brain [52].

Altered expression of zinc transporters in AD brains

There is increasing evidence that zinc transporters are connected with the formation of senile plaques in AD. Abundant expression and altered distribution of zinc transporters have been found in the senile plaques of APP/PS1 mice and human AD brains [56, 59]. ZnT1 and ZnT4 were present in the entire body of the senile plaques, ZnT3, ZnT5 and ZnT6 existed in the peripheral parts of the plaque, while ZnT7 was expressed in the center of the senile plaques [59]. Strikingly, the expression level of zinc transporters also changed abnormally during the progression of AD, including six ZnT transporters and one ZIP transporter (Table 1). Figure 2 shows a schematic depiction of the major altered expression of zinc transporters in the neuronal development of AD.

Fig. 2
figure 2

Schematic depiction of the major altered expression of zinc transporters in the neurons of developing AD. Exchangeable zinc ions across the blood-brain barrier/blood-CSF barrier by binding with His or Cys to form Zn (His)2 or Zn (Cys)(His), and then the complex is transferred into or out of the glial cells and neurons through zinc-binding proteins (ZIPs, ZnTs and DMT1). However, in AD patients, the expression levels of some major zinc transporters are altered, and this exacerbates Aβ deposition and toxicity. As shown in the figure, the highly upregulated ZnT1 pumps more zinc from presynaptic neurons and glial cells, which aggravates the deposition of Aβ proteins, and with the lower available zinc ions in the neuronal cytoplasm, upregulation of ZIP1 expression is induced to import zinc from the extracellular milieu to sustain the normal zinc homeostasis. However, this leads to a vicious cycle. In addition, the decreased expression level of ZnT3 leads to insufficient release of zinc into the cleft, and thus the inhibitory function of zinc on NMDAR will be impaired; as a result, more Ca2+ enters the postsynaptic cells, leading to apoptosis and cognitive disorders

Generally, AD can be classified into four stages: preclinical AD (PCAD), mild cognitive impairment (MCI), early-stage AD (EAD) and late-stage AD (LAD) [79, 80]. MCI subjects have memory loss and a proven AD pathology but with normal daily activities, which usually fall between persons with little or no pathology and persons with AD, while PCAD subjects have evidence of AD neuropathology at autopsy but with normal cognitive functions [2, 80]. The T-tau or P-tau /Aβ42 ratio can be used to predict the progression of MCI to AD with high specificity [81]. The researchers showed that in PCAD patients, the expression level of ZnT1 in the hippocampus was low, accompanied by the increased zinc concentration in the cytoplasm [53]. Whereas, in patients with EAD and LAD, ZnT1 is elevated in the hippocampus gyrus [20]. Meanwhile, other researchers have also substantiated the elevated ZnT1 level in the brain of APP/PS1 mice [59]. Not only ZnT1, but also other ZnT proteins, such as ZnT4 and ZnT6, are additionally implicated in AD. ZnT4 is detected in the lysosomal and endosteum compartments in the hippocampal, and functions by loading cytoplasmic zinc in secretory vesicles [53]. ZnT6 is located in the membrane of the Golgi apparatus, where zinc binds to APP and inhibits the cleavage of APP at the α-secretase site [7, 57]. In the brains of PCAD patients, the level of ZnT1 decreases, concomitant with the significant elevation of ZnT6. In patients with EAD and LAD, both ZnT4 and ZnT6 levels are increased [55]. It can be interpreted that during the progression of AD, the increased level of ZnT1 leads to a higher zinc concentration in the extracellular space, aggravating the deposition of Aβ. On the other hand, when the expression of ZnT6 is higher, more zinc is brought into the trans-Golgi network (TGN), exacerbating Aβ aggregation through inhibition of APP cleavage by α-secretase. However, whether the perturbation of the abundance of ZnT1 affects NMDA receptor function and synaptic plasticity, and how ZnT4 influences the pathogenesis of AD remains unclear.

Some researchers found that the cortical ZnT3 level reduced with aging in wild mice and healthy humans, especially for AD patients the reduction was exacerbated [19]. 6-month-old ZnT3 knockout mice showed defects in learning and memory [18]. Similarly, in the brains of AD patients, the level of ZnT3 is significantly reduced [19]. The researchers proposed that the altered expression level of ZnT3 was due to the disturbance of zinc in synaptic vesicles, with an inadequate release of zinc into the synaptic cleft. Therefore, the inhibitory function of zinc on postsynaptic receptors is impaired, leading to cognitive deficits in AD [18, 82].

In AD patients, the mRNA level of ZnT10 decreases in the frontal cortex, which is parallel to the case in APP/PS1 mice [60]. We can speculate that the dysfunction of ZnT10 contributes to Aβ deposition and senile plaque formation, but whether it is the result of zinc or manganese transported by altered ZnT10 or both remains unclear.

ZIP1 is upregulated in the aging of human frontal cortex [52]. There is evidence that ZIP1 is associated with AD. In the brains of AD patients, the level of ZIP1 is significantly increased with disease progression and Braak staging [61]. This corresponds to the results obtained from the AD model of flies. In Drosophila, there also exist ZIPs and ZnTs. dZIP1 in flies is an ortholog of human ZIP1, which is enriched in the gut and responsible for zinc absorption [83]. Moreover, dZIP1 is also expressed in the brains of flies, and its expression level also rises with aging in Aβ42 flies [84]. These findings suggest that in the progression of AD, zinc interacts more and more with the Aβ and tau proteins, resulting in reduced availability of zinc in the cytoplasm. In order to maintain normal zinc homeostasis, ZIP1 is highly upregulated to uptake zinc from the extracellular milieu.

Altered expression of metallothioneins in AD brains

The main function of metallothioneins is to transport, storage and regulate the cellular essential metals, such as zinc, copper, and iron. MT-3 plays an important role in brain zinc and copper homeostasis, which is expressed mainly in the brain, where it is particularly abundant in zinc-enriched neurons (ZEN), astrocytes, cerebellar cortex, and the early degenerated hippocampus in AD [85, 86]; In contrast, brain MT-1 and MT-2 are mainly expressed in glia cells and mainly in the astrocytes [85]. Considerable work has shown that MT is also abnormally expressed in AD brains [86,87,88], although some conflicting results have been reported. By using a large number of AD cases, it was proved that MT-3 expression is reduced in AD cases at mRNA level and at the protein level in temporal cortex, similar to those for Zn with regional redistribution [87]. In AD mouse, it is also found that MT-3 protein level is reduced in whole brain extracts [89]. However, MT-1 and MT-2 were reported to be up-regulated in AD cases [90,91,92]. It is proposed that the reduced MT-3 contribute to the loss of its protective effects and/or repair functions, leading to an exacerbation of the pathogenic processes, and the up-regulation of MT-1 and MT-2 may be a host defense response to elevated free intracellular Zn2+ levels, reflecting the pathology and inflammatory signals, and that all these changes imply a close relationship between MT and AD.

Novel therapeutic targets for AD

Clioquinol (CQ or PBT1) and the 8-hydroxy quinoline analog PBT2 are two well-known metal ionophores that were developed to treat AD. They could act as chelators to bind excess metals in the brain, such as copper and zinc, as well as metal-protein attenuating compounds (MPACs), thus redistributing the two cations from enriched to depleted brain regions, indicating benefits for ameliorating AD both in mice AD models and in some clinical trials [9,10,11, 37, 93]. However, due to its depletion or redistribution of metal elements throughout the organism, the systemic administration of chelators may cause harmful side effects and create safety problems. PBT1 has been reported to cause myelinopathies in the CNS of a transgenic AD mouse model and wild-type mice [94]. Because of doubts about positive clinical benefits for AD patients, and misgivings about the safety of the drug [95], as well as problems with contamination during the manufacturing process [37], the planned phase III clinical trial of PBT1 has been abandoned, and the compound has been withdrawn. The metal chaperone of PBT2, as an improved chelator, could rapidly decrease soluble interstitial brain Aβ in a matter of hours and ameliorate the cognitive performance of AD mice [11]. Although PBT2 appeared to be safer, and resulted in a reduced amount of soluble plaque in Phase II trials in Europe (EURO) and Australia (IMAGINE), cognitive outcomes were different and restricted to two measures of executive function [96, 97]. Larger trials are needed to demonstrate its cognitive efficacy [98].

Since zinc transporters are critically involved in brain zinc homeostasis and AD pathological processes, it is possible to specifically modulate the expression level of brain zinc transporter to produce significant ameliorating effects on the progress of AD (as illustrated in Fig. 3). In the Drosophila AD model, Lang et al. [48] reported that dZIP1 knockdown of Aβ42 flies led to a longer lifespan (Fig. 3d), a stronger climbing ability and fewer brain vacuoles than the control of Aβ42 flies, whereas dZIP1 overexpression exacerbated the AD-like symptoms of Aβ42 flies. Furthermore, this study suggested an internal mechanism since the ameliorated AD-like symptoms were well correlated with significantly reduced Aβ42 fibril deposits (Fig. 3a-c) and its SDS-soluble forms, as well as reduced levels of zinc accumulation in brains of aged AD flies upon dZIP1 inhibition [84]. Recent cases reported by Huang et al. [99] also supports the hypothesis in another Drosophila AD model of tauopathy flies, which expresses a human Tau mutant (hTauR406W, or Tau*). Either knocking down dZIP1 or overexpressing dZnT1 significantly reduced the number of brain vacuoles and prolonged the lifespan of Tau* flies, demonstrating the ameliorating effects of zinc limitation through the genetic modulation of brain zinc transporters, where Tau phosphorylation and the toxicity produced by zinc-Tau binding were markedly reduced [99]. Another study in mice showed that knockout of ZIP1 and ZIP3 can attenuate seizure-induced neurodegeneration in hippocampal CA1 pyramidal neurons [62]. Moreover, 6 month-old ZnT3 knockout mice showed deficits in learning and memory [18, 68], and a further study showed that ZnT3 levels were negatively correlated with soluble Aβ [100], while the ablation of ZnT3 markedly reduced the plaque burden in hAPP+:ZnT3−/− mice [101]. Roy et al. (2017) reported the design of compounds aimed at preventing MT-3 degradation in AD cases, and showed that Benzothiazolone-2 compounds significantly enhanced MT-3 protein and mRNA levels, suggesting its use as a lead compound for future drug design in the treatment of AD [102]. The surprising efficiency of the influence on AD-like symptoms in animal models, when modulating zinc transporters, suggests a novel strategy for the treatment of AD. In fact, correcting zinc homeostasis in the brains of AD patients by altering the expression level or traffic capacity of zinc transporters holds promise as an effective therapeutic strategy that should be further explored.

Fig. 3
figure 3

Modifying dZIP1 levels markedly influences the brain Aβ42 fibril deposition and lifespan in a Drosophila AD model. a-b Thioflavin-S (TS) staining was used to detect Aβ42 fibril deposits (bright green dots) in fly brains. Few deposits were found in the control brains (Elav-Gal4, top left) at 25 days after eclosion (dae). TS-positive deposits were found after Aβ42 expression in fly brains (Elav-Gal4 > UAS-Aβ42) at both a 25 and b 30 dae. c The quantitative content of Aβ42 deposits was summarized and expressed after normalization to 25-day old Aβ42 flies. The increase in Aβ42 deposits was age-dependent. Overexpression of dZIP1 in Aβ42-expressing brains (Elav-Gal4 > UAS-Aβ42/UAS-dZIP1) significantly increased fibril deposits at 25 dae, which was higher than 30 dae Aβ42 flies. However, inhibition of dZIP1 (Elav-Gal4 > UAS-Aβ42/UAS-dZIP1 RNAi) dramatically decreased deposit density at 30 dae, which was reduced compared to 25 dae Aβ42 flies. t test, **P < 0.01, ***P < 0.001. Data are expressed as means ± SEM. n = 6 or 8 hemispheres for each genotype. Scale bar: 25 μm. d dZip1 knockdown significantly prolongs the lifespan of Aβ42 flies. The percentage of survivorship was plotted against age (dae). Overexpression of dZIP1significantly shortened the lifespan of Aβ42 (elav-Gal4 > UAS-Aβ42) flies. Decreased dZIP1 levels (Elav-Gal4 > UAS-Aβ42/UAS-dZip1 RNAi) inhibited Aβ42 toxicity in a dose-dependent manner, and dZip1 RNAi #2 showed a more significant phenotype, in which dZip1-RNAi 2# (elav-Gal4 > UAS-Aβ42/UAS-dZip1-RNAi 2#) flies had much more reduced dZIP1 level than that of dZip1-RNAi 1# (elav-Gal4 > UAS-Aβ42/UAS-dZip1-RNAi 1#) flies. The differences shown are all statistically significant (p < 0.001). The reported P values are derived from Mantel-Cox log-rank statistical analysis

Future questions and perspectives

In this review, we highlight the significant role of zinc transporters in the pathology of AD and catalogue their multiple pathological changes in disease progression. Figure 4 is a proposed model that illustrates the mechanisms of AD progression and amelioration strategy; however, some questions remain to be resolved. How do zinc transporters deliver zinc across the blood-brain barrier? Why do the expression levels of ZnT1, ZnT3, ZnT4, ZnT6, ZnT10 and Zip1 alter at different stages of AD? Furthermore, the way ZnTs and ZIPs exist in the glia to modulate zinc homeostasis in AD are also critical issues that must be resolved. We need to use AD models, such as Drosophila, to uncover how ZnTs and ZIPs interact to influence the development of AD pathology, and utilize the methods of gene therapy or the development of lead compounds to alter the expression or activity of zinc transporters in mice or higher animal models to investigate whether AD-like symptoms can be prevented or cured.

Fig. 4
figure 4

A proposed model illustrating the mechanism of AD progression and amelioration. The expression or alteration in activity of brain zinc transporters induce zinc dyshomeostasis, which aggravate Aß deposition, tau phosphorylation and tau-zinc binding exacerbated toxicity, increasingly promoting neuronal loss (a), while the leading compounds or drugs designed to specifically modify the expression or activity of brain zinc transport proteins have the potential to correct the disturbed zinc metabolism niche, thus leading to reduced Aß deposition, tau dissociation from microtubules and tau toxicity, which ultimately slows or prevents neuronal death in the onset and progression of AD (b)

If such experiments are successful, in subsequent clinical trials, we need to find appropriate delivery methods for human-specific brain regions to specifically target the transporter mRNAs or proteins responsible for enhancing or weakening zinc levels or activity, in order to correct the disturbing zinc metabolism niche. Importantly, these genetic methods or novel pharmacological molecules used to reform the activity of zinc transporters may allow the discovery of novel therapies not only for AD, but also for other neurodegenerative diseases such as PD and HD. In summary, further exploration of the potential of transporter therapeutic targets may shed new light on the prevention or treatment of chronic neurodegenerative diseases.

Availability of data and materials

Not applicable.

Abbreviations

AD:

Alzheimer’s disease

AE:

Acrodermatitis enteropathica

AMPA:

α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate

APP:

Amyloid precursor protein

Aβ:

Amyloid β

BDNF:

Brain-derived neurotrophic factor

CNS:

Central nervous system

CQ:

Clioquinol

EAD:

Early-stage AD

ER:

Endoplasmic reticulum

HD:

Huntington’s disease

His:

Histidine

LAD:

Late-stage AD

mAR:

membrane androgen receptor

MPP+:

1-Methyl-4-phenylpyridinium.

MTS:

Metallothioneins

NFTs:

Neurofibrillary tangles

NMDA:

N-methyl-D-aspartate

PBT2:

5,7-Dichloro-2-[(dimethylamino)methyl]quinolin-8-ol

PCAD:

Preclinical AD

PD:

Parkinson’s disease

PS1:

Presenilin 1

SCD-EDS:

Ehlers–Danlos syndrome

SLC30:

Solute carrier 30

SLC39:

Solute carrier 39

TGN:

Trans-Golgi network

TMDs:

Transmembrane domains

TNAP:

Tissue-nonspecific alkaline phosphatase

TrkB:

Tropomyosin-related kinase B

VDCC:

Voltage-dependent calcium channels

ZIP:

Zrt- and Irt-like protein

ZnT:

Zinc transporter

References

  1. Calap-Quintana P, Gonzalez-Fernandez J, Sebastia-Ortega N, Llorens JV, Molto MD. Drosophila melanogaster models of metal-related human diseases and metal toxicity. Int J Mol Sci. 2017;18(7). https://doi.org/10.3390/ijms18071456.

    Article  Google Scholar 

  2. Winblad B, Amouyel P, Andrieu S, Ballard C, Brayne C, Brodaty H, Cedazo-Minguez A, Dubois B, Edvardsson D, Feldman H, Fratiglioni L, Frisoni GB, Gauthier S, Georges J, Graff C, Iqbal K, Jessen F, Johansson G, Jonsson L, Kivipelto M, Knapp M, Mangialasche F, Melis R, Nordberg A, Rikkert MO, Qiu C, Sakmar TP, Scheltens P, Schneider LS, Sperling R, Tjernberg LO, Waldemar G, Wimo A, Zetterberg H. Defeating Alzheimer's disease and other dementias: a priority for European science and society. Lancet Neurol. 2016;15(5):455–532. https://doi.org/10.1016/S1474-4422(16)00062-4.

    Article  PubMed  Google Scholar 

  3. Overk CR, Masliah E. Pathogenesis of synaptic degeneration in Alzheimer's disease and Lewy body disease. Biochem Pharmacol. 2014;88(4):508–16. https://doi.org/10.1016/j.bcp.2014.01.015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Selkoe DJ. The molecular pathology of Alzheimers-disease. Neuron. 1991;6(4):487–98. https://doi.org/10.1016/0896-6273(91)90052-2.

    Article  CAS  PubMed  Google Scholar 

  5. Savelieff MG, Lee S, Liu Y, Lim MH. Untangling amyloid-beta, tau, and metals in Alzheimer's disease. ACS Chem Biol. 2013;8(5):856–65. https://doi.org/10.1021/cb400080f.

    Article  CAS  PubMed  Google Scholar 

  6. Bush AI, Pettingell WH, Paradis MD, Tanzi RE. Modulation of a-BETA adhesiveness and SECRETASE site cleavage by zinc. J Biol Chem. 1994;269(16):12152–8.

    CAS  PubMed  Google Scholar 

  7. Bush AI, Pettingell WH, Multhaup G, Paradis MD, Vonsattel JP, Gusella JF, Beyreuther K, Masters CL, Tanzi RE. Rapid induction of Alzheimer a-beta amyloid formation by zinc. Science. 1994;265(5177):1464–7. https://doi.org/10.1126/science.8073293.

    Article  CAS  Google Scholar 

  8. Kim I, Park EJ, Seo J, Ko SJ, Lee J, Kim CH. Zinc stimulates tau S214 phosphorylation by the activation of Raf/mitogen-activated protein kinase-kinase/extracellular signal-regulated kinase pathway. Neuroreport. 2011;22(16):839–44. https://doi.org/10.1097/WNR.0b013e32834c0a2d.

    Article  CAS  PubMed  Google Scholar 

  9. Ritchie CW, Bush AI, Mackinnon A, Macfarlane S, Mastwyk M, MacGregor L, Kiers L, Cherny R, Li QX, Tammer A, Carrington D, Mavros C, Volitakis I, Xilinas M, Ames D, Davis S, Volitakis I, Xilinas M, Ames D, Davis S, Beyreuther K, Tanzi RE, Masters CL. Metal-protein attenuation with iodochlorhydroxyquin (clioquinol) targeting a beta amyloid deposition and toxicity in Alzheimer disease - a pilot phase 2 clinical trial. Arch Neurol. 2003;60(12):1685–91. https://doi.org/10.1001/archneur.60.12.1685.

    Article  PubMed  Google Scholar 

  10. Lannfelt L, Blennow K, Zetterberg H, Batsman S, Ames D, Hrrison J, Masters CL, Targum S, Bush AI, Murdoch R, Wilson J, Ritchie CW, Grp PES. Safety, efficacy, and biomarker findings of PBT2 in targeting a beta as a modifying therapy for Alzheimer's disease: a phase IIa, double-blind, randomised, placebo-controlled trial. Lancet Neurol. 2008;7(9):779–86. https://doi.org/10.1016/s1474-4422(08)70167-4.

    Article  CAS  Google Scholar 

  11. Adlard PA, Cherny RA, Finkelstein DI, Gautier E, Robb E, Cortes M, Volitakis I, Liu X, Smith JP, Perez K, Laughton K, Li QX, Charman SA, Nicolazzo JA, Wilkins S, Deleva K, Lynch T, Kok G, Ritchie CW, Tanzi RE, Cappai R, Masters CL, Barnham KJ, Bush AI. Rapid restoration of cognition in Alzheimer's transgenic mice with 8-hydroxy quinoline analogs is associated with decreased interstitial a beta. Neuron. 2008;59(1):43–55. https://doi.org/10.1016/j.neuron.2008.06.018.

    Article  CAS  PubMed  Google Scholar 

  12. Sindreu C, Bayes A, Altafaj X, Perez-Clausell J. Zinc transporter-1 concentrates at the postsynaptic density of hippocampal synapses. Mol Brain. 2014;7. https://doi.org/10.1186/1756-6606-7-16.

    Article  Google Scholar 

  13. Smidt K, Rungby J. ZnT3: a zinc transporter active in several organs. Biometals. 2012;25(1):1–8. https://doi.org/10.1007/s10534-011-9490-x.

    Article  CAS  PubMed  Google Scholar 

  14. Lichten LA, Cousins RJ. Mammalian zinc transporters: nutritional and physiologic regulation. Annu Rev Nutr. 2009;29:153–76. https://doi.org/10.1146/annurev-nutr-033009-083312.

    Article  PubMed  Google Scholar 

  15. Belloni-Olivi L, Marshall C, Laal B, Andrews GK, Bressler J. Localization of Zip1 and Zip4 mRNA in the adult rat brain. J Neurosci Res. 2009;87(14):3221–30. https://doi.org/10.1002/jnr.22144.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Segawa S, Tatsumi N, Ohishi A, Nishida K, Nagasawa K. Characterization of zinc uptake by mouse primary cultured astrocytes and microglia. Metallomics. 2015;7(7):1067–77. https://doi.org/10.1039/c5mt00085h.

    Article  CAS  PubMed  Google Scholar 

  17. Chowanadisai W, Kelleher SL, Lonnerdal B. Zinc deficiency is associated with increased brain zinc import and LIV-1 expression and decreased ZnT-1 expression in neonata rats. J Nutr. 2005;135(5):1002–7.

    Article  CAS  Google Scholar 

  18. Adlard PA, Parncutt JM, Finkelstein DI, Bush AI. 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(5):1631–6. https://doi.org/10.1523/JNEUROSCI.5255-09.2010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Beyer N, Coulson DT, Heggarty S, Ravid R, Irvine GB, Hellemans J, Johnston JA. ZnT3 mRNA levels are reduced in Alzheimer's disease post-mortem brain. Mol Neurodegener. 2009;4:53. https://doi.org/10.1186/1750-1326-4-53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lovell MA, Smith JL, Xiong S, Markesbery WR. Alterations in zinc transporter protein-1 (ZnT-1) in the brain of subjects with mild cognitive impairment, early, and late-stage Alzheimer's disease. Neurotox Res. 2005;7(4):265–71.

    Article  CAS  Google Scholar 

  21. Oteiza PI, Mackenzie GG. Zinc, oxidant-triggered cell signaling, and human health. Mol Asp Med. 2005;26(4–5):245–55. https://doi.org/10.1016/j.mam.2005.07.012.

    Article  CAS  Google Scholar 

  22. Barceloux DG. Zinc. J Toxicol Clin Toxicol. 1999;37(2):279–92.

    Article  CAS  Google Scholar 

  23. Nakashima AS, Dyck RH. Zinc and cortical plasticity. Brain Res Rev. 2009;59(2):347–73. https://doi.org/10.1016/j.brainresrev.2008.10.003.

    Article  CAS  Google Scholar 

  24. Plum LM, Rink L, Haase H. The essential toxin: impact of zinc on human health. Int J Environ Res Public Health. 2010;7(4):1342–65. https://doi.org/10.3390/ijerph7041342.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Bitanihirwe BK, Cunningham MG. Zinc: the brain's dark horse. Synapse (New York, NY). 2009;63(11):1029–49. https://doi.org/10.1002/syn.20683.

    Article  CAS  Google Scholar 

  26. Frederickson CJ. Neurobiology of zinc and zinc-containing neurons. Int Rev Neurobiol. 1989;31:145–238. https://doi.org/10.1016/s0074-7742(08)60279-2.

    Article  CAS  PubMed  Google Scholar 

  27. Slomianka L. Neurons of origin of zinc-containing pathways and the distribution of zinc-containing BOUTONS in the hippocampal region of the rat. Neuroscience. 1992;48(2):325–52. https://doi.org/10.1016/0306-4522(92)90494-m.

    Article  CAS  PubMed  Google Scholar 

  28. Beaulieu C, Dyck R, Cynader M. Enrichment of glutamate in zinc-containing terminals of the cat visual-cortex. Neuroreport. 1992;3(10):861–4. https://doi.org/10.1097/00001756-199210000-00010.

    Article  CAS  PubMed  Google Scholar 

  29. Sensi SL, Yin HZ, Carriedo SG, Rao SS, Weiss JH. Preferential Zn2+ influx through Ca2+−permeable AMPA/kainate channels triggers prolonged mitochondrial superoxide production. Proc Natl Acad Sci U S A. 1999;96(5):2414–9.

    Article  CAS  Google Scholar 

  30. Kerchner GA, Canzoniero LM, Yu SP, Ling C, Choi DW. Zn2+ current is mediated by voltage-gated Ca2+ channels and enhanced by extracellular acidity in mouse cortical neurones. J Physiol. 2000;528(Pt 1):39–52. https://doi.org/10.1111/j.1469-7793.2000.00039.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Koh JY, Choi DW. Zinc toxicity on cultured cortical neurons: involvement of N-methyl-D-aspartate receptors. Neuroscience. 1994;60(4):1049–57.

    Article  CAS  Google Scholar 

  32. Karakas E, Simorowski N, Furukawa H. Structure of the zinc-bound amino-terminal domain of the NMDA receptor NR2B subunit. EMBO J. 2009;28(24):3910–20. https://doi.org/10.1038/emboj.2009.338.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Marin P, Israel M, Glowinski J, Premont J. Routes of zinc entry in mouse cortical neurons: role in zinc-induced neurotoxicity. Eur J Neurosci. 2000;12(1):8–18. https://doi.org/10.1046/j.1460-9568.2000.00875.x.

    Article  CAS  PubMed  Google Scholar 

  34. Sensi SL, Jeng JM. Rethinking the excitotoxic ionic milieu: the emerging role of Zn(2+) in ischemic neuronal injury. Curr Mol Med. 2004;4(2):87–111. https://doi.org/10.2174/1566524043479211.

    Article  CAS  PubMed  Google Scholar 

  35. Dineley KE, Votyakova TV, Reynolds IJ. Zinc inhibition of cellular energy production: implications for mitochondria and neurodegeneration. J Neurochem. 2003;85(3):563–70. https://doi.org/10.1046/j.1471-4159.2003.01678.x.

    Article  CAS  PubMed  Google Scholar 

  36. Gazaryan IG, Krasinskaya IP, Kristal BS, Brown AM. Zinc irreversibly damages major enzymes of energy production and antioxidant defense prior to mitochondrial permeability transition. J Biol Chem. 2007;282(33):24373–80. https://doi.org/10.1074/jbc.M611376200.

    Article  CAS  PubMed  Google Scholar 

  37. Barnham KJ, Bush AI. Biological metals and metal-targeting compounds in major neurodegenerative diseases. Chem Soc Rev. 2014;43(19):6727–49. https://doi.org/10.1039/c4cs00138a.

    Article  CAS  PubMed  Google Scholar 

  38. Mao Y, Zang S, Zhang J, Sun X. Early chronic blockade of NR2B subunits and transient activation of NMDA receptors modulate LTP in mouse auditory cortex. Brain Res. 2006;1073-1074:131–8. https://doi.org/10.1016/j.brainres.2005.12.077.

    Article  CAS  PubMed  Google Scholar 

  39. Takeda A, Suzuki M, Tamano H, Ando M, Oku N. Differential effects of zinc influx via AMPA/kainate receptor activation on subsequent induction of hippocampal CA1 LTP components. Brain Res. 2010;1354:188–95. https://doi.org/10.1016/j.brainres.2010.07.039.

    Article  CAS  PubMed  Google Scholar 

  40. Huang YZ, Pan E, Xiong ZQ, McNamara JO. Zinc-mediated transactivation of TrkB potentiates the hippocampal mossy fiber-CA3 pyramid synapse. Neuron. 2008;57(4):546–58. https://doi.org/10.1016/j.neuron.2007.11.026.

    Article  CAS  PubMed  Google Scholar 

  41. Corona C, Pensalfini A, Frazzini V, Sensi SL. New therapeutic targets in Alzheimer's disease: brain deregulation of calcium and zinc. Cell Death Dis. 2011;2:e176. https://doi.org/10.1038/cddis.2011.57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Corona C, Masciopinto F, Silvestri E, Viscovo AD, Lattanzio R, Sorda RL, Ciavardelli D, Goglia F, Piantelli M, Canzoniero LM, Sensi SL. Dietary zinc supplementation of 3xTg-AD mice increases BDNF levels and prevents cognitive deficits as well as mitochondrial dysfunction. Cell Death Dis. 2010;1:e91. https://doi.org/10.1038/cddis.2010.73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Helgager J, Huang YZ, McNamara JO. Brain-derived neurotrophic factor but not vesicular zinc promotes TrkB activation within mossy fibers of mouse hippocampus in vivo. J Comp Neurol. 2014;522(17):3885–99. https://doi.org/10.1002/cne.23647.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Abramovitch-Dahan C, Asraf H, Bogdanovic M, Sekler I, Bush AI, Hershfinkel M. Amyloid β attenuates metabotropic zinc sensing receptor, mZnR/GPR39, dependent Ca2+, ERK1/2 and Clusterin signaling in neurons. J Neurochem. 2016;139(2):221–33. https://doi.org/10.1111/jnc.13760.

    Article  CAS  PubMed  Google Scholar 

  45. Xiao G, Zhou B. What can flies tell us about zinc homeostasis? Arch Biochem Biophys. 2016;611:134–41. https://doi.org/10.1016/j.abb.2016.04.016.

    Article  CAS  PubMed  Google Scholar 

  46. Coyle P, Philcox JC, Carey LC, Rofe AM. Metallothionein: the multipurpose protein. Cell Mol Life Sci. 2002;59(4):627–47. https://doi.org/10.1007/s00018-002-8454-2.

    Article  CAS  PubMed  Google Scholar 

  47. Kambe T, Hashimoto A, Fujimoto S. Current understanding of ZIP and ZnT zinc transporters in human health and diseases. Cell Mol Life Sci. 2014;71(17):3281–95. https://doi.org/10.1007/s00018-014-1617-0.

    Article  CAS  PubMed  Google Scholar 

  48. Fukunaka A, Suzuki T, Kurokawa Y, Yamazaki T, Fujiwara N, Ishihara K, Migaki H, Okumura K, Masuda S, Yamaguchi-Iwai Y, Nagao M, Kambe T. Demonstration and characterization of the Heterodimerization of ZnT5 and ZnT6 in the early secretory pathway. J Biol Chem. 2009;284(45):30798–806. https://doi.org/10.1074/jbc.M109.026435.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Quadri M, Federico A, Zhao T, Breedveld GJ, Battisti C, Delnooz C, Severijnen L-A, Mammarella LDT, Mignarri A, Monti L, Sanna A, Lu P, Punzo F, Cossu G, Willemsen R, Rasi F, Oostra BA, van de Warrenburg BP, Bonifati V. Mutations in SLC30A10 cause parkinsonism and dystonia with Hypermanganesemia, polycythemia, and chronic liver disease. Am J Hum Genet. 2012;90(3):467–77. https://doi.org/10.1016/j.ajhg.2012.01.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Braidy N, Poljak A, Marjo C, Rutlidge H, Rich A, Jugder B-E, Jayasena T, Inestrosa NC, Sachdev PS. Identification of cerebral metal ion imbalance in the brain of aging Octodon degus. Front Aging Neurosci. 2017;9(66). https://doi.org/10.3389/fnagi.2017.00066.

  51. Adlard PA, Parncutt J, Lal V, James S, Hare D, Doble P, Finkelstein DI, Bush AI. Metal chaperones prevent zinc-mediated cognitive decline. Neurobiol Dis. 2015;81:196–202. https://doi.org/10.1016/j.nbd.2014.12.012.

    Article  CAS  PubMed  Google Scholar 

  52. Olesen RH, Hyde TM, Kleinman JE, Smidt K, Rungby J, Larsen A. Obesity and age-related alterations in the gene expression of zinc-transporter proteins in the human brain. Transl Psychiatry. 2016;6(6):e838. https://doi.org/10.1038/tp.2016.83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Lyubartseva G, Smith JL, Markesbery WR, Lovell MA. Alterations of zinc transporter proteins ZnT-1, ZnT-4 and ZnT-6 in preclinical Alzheimer's disease brain. Brain Pathol. 2010;20(2):343–50. https://doi.org/10.1111/j.1750-3639.2009.00283.x.

    Article  CAS  PubMed  Google Scholar 

  54. Palmiter RD, Cole TB, Quaife CJ, Findley SD. ZnT-3, a putative transporter of zinc into synaptic vesicles. Proc Natl Acad Sci U S A. 1996;93(25):14934–9. https://doi.org/10.1073/pnas.93.25.14934.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Smith JL, Xiong S, Markesbery WR, Lovell MA. Altered expression of zinc transporters-4 and -6 in mild cognitive impairment, early and late Alzheimer's disease brain. Neuroscience. 2006;140(3):879–88. https://doi.org/10.1016/j.neuroscience.2006.02.049.

    Article  CAS  PubMed  Google Scholar 

  56. Zhang LH, Wang X, Stoltenberg M, Danscher G, Huang L, Wang ZY. Abundant expression of zinc transporters in the amyloid plaques of Alzheimer's disease brain. Brain Res Bull. 2008;77(1):55–60. https://doi.org/10.1016/j.brainresbull.2008.03.014.

    Article  CAS  PubMed  Google Scholar 

  57. Baulac S, LaVoie MJ, Kimberly WT, Strahle J, Wolfe MS, Selkoe DJ, Xia WM. Functional gamma-secretase complex assembly in Golgi/trans-Golgi network: interactions among presenilin, nicastrin, Aph1, Pen-2, and gamma-secretase substrates. Neurobiol Dis. 2003;14(2):194–204. https://doi.org/10.1016/s0969-9961(03)00123-2.

    Article  CAS  PubMed  Google Scholar 

  58. Kirschke CP, Huang LP. ZnT7, a novel mammalian zinc transporter, accumulates zinc in the Golgi apparatus. J Biol Chem. 2003;278(6):4096–102. https://doi.org/10.1074/jbc.M207644200.

    Article  CAS  PubMed  Google Scholar 

  59. Zhang LH, Wang X, Zheng ZH, Ren H, Stoltenberg M, Danscher G, Huang L, Rong M, Wang ZY. Altered expression and distribution of zinc transporters in APP/PS1 transgenic mouse brain. Neurobiol Aging. 2010;31(1):74–87. https://doi.org/10.1016/j.neurobiolaging.2008.02.018.

    Article  CAS  PubMed  Google Scholar 

  60. Bosomworth HJ, Adlard PA, Ford D, Valentine RA. Altered expression of ZnT10 in Alzheimer's disease brain. PLoS One. 2013;8(5). https://doi.org/10.1371/journal.pone.0065475.

    Article  CAS  Google Scholar 

  61. Beyer N, Coulson DT, Heggarty S, Ravid R, Hellemans J, Irvine GB, Johnston JA. Zinc transporter mRNA levels in Alzheimer's disease postmortem brain. J Alzheimers Dis. 2012;29(4):863–73. https://doi.org/10.3233/JAD-2012-112105.

    Article  CAS  PubMed  Google Scholar 

  62. Qian J, Xu K, Yoo J, Chen TT, Andrews G, Noebels JL. Knockout of Zn transporters zip-1 and zip-3 attenuates seizure-induced CA1 Neurodegeneration. J Neurosci. 2011;31(1):97–104. https://doi.org/10.1523/jneurosci.5162-10.2011.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Lin Y, Chen Y, Wang YZ, Yang JX, Zhu VF, Liu YL, Cui XB, Chen L, Yan W, Jiang T, Hergenroeder GW, Fletcher SA, Levine JM, Kim DH, Tandon N, Zhu JJ, Li M. ZIP4 is a novel molecular marker for glioma. Neuro-Oncology. 2013;15(8):1008–16. https://doi.org/10.1093/neuonc/not042.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Matsuura W, Yamazaki T, Yamaguchi-Iwai Y, Masuda S, Nagao M, Andrews GK, Kambe T. SLC39A9 (ZIP9) regulates zinc homeostasis in the secretory pathway: characterization of the ZIP subfamily I protein in vertebrate cells. Biosci Biotechnol Biochem. 2009;73(5):1142–8. https://doi.org/10.1271/bbb.80910.

    Article  CAS  PubMed  Google Scholar 

  65. Bin B-H, Hojyo S, Ryong Lee T, Fukada T. Spondylocheirodysplastic Ehlers-Danlos syndrome (SCD-EDS) and the mutant zinc transporter ZIP13. Rare Dis (Austin, Tex). 2014;2(1):e974982. https://doi.org/10.4161/21675511.2014.974982.

    Article  Google Scholar 

  66. McMahon RJ, Cousins RJ. Regulation of the zinc transporter ZnT-1 by dietary zinc. Proc Natl Acad Sci U S A. 1998;95(9):4841–6. https://doi.org/10.1073/pnas.95.9.4841.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Mellone M, Pelucchi S, Alberti L, Geneazzani AA, Di Luca M, Gardoni F. Zinc transporter-1: a novel NMDA receptor-binding protein at the postsynaptic density. J Neurochem. 2015;132(2):159–68. https://doi.org/10.1111/jnc.12968.

    Article  CAS  PubMed  Google Scholar 

  68. Cole TB, Wenzel HJ, Kafer KE, Schwartzkroin PA, Palmiter RD. Elimination of zinc from synaptic vesicles in the intact mouse brain by disruption of the ZnT3 gene. Proc Natl Acad Sci U S A. 1999;96(4):1716–21. https://doi.org/10.1073/pnas.96.4.1716.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Huang LP, Gitschier J. A novel gene involved in zinc transport is deficient in the lethal milk mouse. Nat Genet. 1997;17(3):292–7. https://doi.org/10.1038/ng1197-292.

    Article  CAS  PubMed  Google Scholar 

  70. Bosomworth HJ, Thornton JK, Coneyworth LJ, Ford D, Valentine RA. Efflux function, tissue-specific expression and intracellular trafficking of the Zn transporter ZnT10 indicate roles in adult Zn homeostasis. Metallomics. 2012;4(8):771–9. https://doi.org/10.1039/c2mt20088k.

    Article  CAS  PubMed  Google Scholar 

  71. Nishito Y, Tsuji N, Fujishiro H, Takeda T, Yamazaki T, Teranishi F, Okazaki F, Matsunaga A, Tuschl K, Rao R, Kono S, Miyajima H, Narita H, Himeno S, Kambe T. Direct comparison of manganese detoxification/efflux proteins and molecular characterization of ZnT10 protein as a manganese transporter. J Biol Chem. 2016;291(28):14773–87. https://doi.org/10.1074/jbc.M116.728014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Fujishiro H, Yoshida M, Nakano Y, Himeno S. Interleukin-6 enhances manganese accumulation in SH-SY5Y cells: implications of the up-regulation of ZIP14 and the down-regulation of ZnT10. Metallomics. 2014;6(4):944–9. https://doi.org/10.1039/c3mt00362k.

    Article  CAS  PubMed  Google Scholar 

  73. Delhaize E, Kataoka T, Hebb DM, White RG, Ryan PR. Genes encoding proteins of the cation diffusion facilitator family that confer manganese tolerance. Plant Cell. 2003;15(5):1131–42.

    Article  CAS  Google Scholar 

  74. Costello LC, Liu YY, Zou J, Franklin RB. Evidence for a zinc uptake transporter in human prostate cancer cells which is regulated by prolactin and testosterone. J Biol Chem. 1999;274(25):17499–504. https://doi.org/10.1074/jbc.274.25.17499.

    Article  CAS  PubMed  Google Scholar 

  75. Gaither LA, Eide DJ. The human ZIP1 transporter mediates zinc uptake in human K562 erythroleukemia cells. J Biol Chem. 2001;276(25):22258–64. https://doi.org/10.1074/jbc.M101772200.

    Article  CAS  PubMed  Google Scholar 

  76. 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(12):1933–45. https://doi.org/10.1002/glia.21235.

    Article  PubMed  Google Scholar 

  77. Thomas P, Pang YF, Dong J, Berg AH. Identification and characterization of membrane androgen receptors in the ZIP9 zinc transporter subfamily: II. Role of human ZIP9 in testosterone-induced prostate and breast Cancer cell apoptosis. Endocrinology. 2014;155(11):4250–65. https://doi.org/10.1210/en.2014-1201.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Xiao GR, Wan ZH, Fan QW, Tang XN, Zhou B. The metal transporter ZIP13 supplies iron into the secretory pathway in Drosophila melanogaster. eLife. 2014;3:21. https://doi.org/10.7554/eLife.03191.

    Article  CAS  Google Scholar 

  79. Triplett JC, Swomley AM, Cai J, Klein JB, Butterfield DA. Quantitative phosphoproteomic analyses of the inferior parietal lobule from three different pathological stages of Alzheimer's disease. J Alzheimers Dis. 2016;49(1):45–62. https://doi.org/10.3233/JAD-150417.

    Article  CAS  PubMed  Google Scholar 

  80. Tramutola A, Triplett JC, Di Domenico F, Niedowicz DM, Murphy MP, Coccia R, Perluigi M, Butterfield DA. Alteration of mTOR signaling occurs early in the progression of Alzheimer disease (AD): analysis of brain from subjects with pre-clinical AD, amnestic mild cognitive impairment and late-stage AD. J Neurochem. 2015;133(5):739–49. https://doi.org/10.1111/jnc.13037.

    Article  CAS  PubMed  Google Scholar 

  81. Dubois B, Feldman HH, Jacova C, Hampel H, Molinuevo JL, Blennow K, DeKosky ST, Gauthier S, Selkoe D, Bateman R, Cappa S, Crutch S, Engelborghs S, Frisoni GB, Fox NC, Galasko D, Habert MO, Jicha GA, Nordberg A, Pasquier F, Rabinovici G, Robert P, Rowe C, Salloway S, Sarazin M, Epelbaum S, de Souza LC, Vellas B, Visser PJ, Schneider L, Stern Y, Scheltens P, Cummings JL. Advancing research diagnostic criteria for Alzheimer's disease: the IWG-2 criteria. Lancet Neurol. 2014;13(6):614–29. https://doi.org/10.1016/s1474-4422(14)70090-0.

    Article  PubMed  Google Scholar 

  82. Deshpande A, Kawai H, Metherate R, Glabe CG, Busciglio J. A role for synaptic zinc in activity-dependent a beta oligomer formation and accumulation at excitatory synapses. J Neurosci. 2009;29(13):4004–15. https://doi.org/10.1523/jneurosci.5980-08.2009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Qin Q, Wang X, Zhou B. Functional studies of Drosophila zinc transporters reveal the mechanism for dietary zinc absorption and regulation. BMC Biol. 2013;11:101. https://doi.org/10.1186/1741-7007-11-101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Lang M, Wang L, Fan Q, Xiao G, Wang X, Zhong Y, Zhou B. Genetic inhibition of solute-linked carrier 39 family transporter 1 ameliorates a beta pathology in a Drosophila model of Alzheimer's disease. PLoS Genet. 2012;8(4):623–39. https://doi.org/10.1371/journal.pgen.1002683.

    Article  CAS  Google Scholar 

  85. Kepp KP. Bioinorganic chemistry of Alzheimer's disease. Chem Rev. 2012;112(10):5193–239. https://doi.org/10.1021/cr300009x.

    Article  CAS  PubMed  Google Scholar 

  86. Carrasco J, Giralt M, Molinero A, Penkowa M, Moos T, Hidalgo J. Metallothionein (MT)-III: generation of polyclonal antibodies, comparison with MT-I+II in the freeze lesioned rat brain and in a bioassay with astrocytes, and analysis of Alzheimer's disease brains. J Neurotrauma. 1999;16(11):1115–29. https://doi.org/10.1089/neu.1999.16.1115.

    Article  CAS  PubMed  Google Scholar 

  87. Yu WH, Lukiw WJ, Bergeron C, Niznik HB, Fraser PE. Metallothionein III is reduced in Alzheimer's disease. Brain Res. 2001;894(1):37–45. https://doi.org/10.1016/s0006-8993(00)03196-6.

    Article  CAS  PubMed  Google Scholar 

  88. Uchida Y, Takio K, Titani K, Ihara Y, Tomonaga M. The growth inhibitory factor that is deficient in the Alzheimer's disease brain is a 68 amino acid metallothionein-like protein. Neuron. 1991;7(2):337–47. https://doi.org/10.1016/0896-6273(91)90272-2.

    Article  CAS  PubMed  Google Scholar 

  89. Martin BL, Tokheim AM, McCarthy PT, Doms BS, Davis AA, Armitage IM. Metallothionein-3 and neuronal nitric oxide synthase levels in brains from the Tg2576 mouse model of Alzheimer's disease. Mol Cell Biochem. 2006;283(1–2):129–37. https://doi.org/10.1007/s11010-006-2390-7.

    Article  CAS  PubMed  Google Scholar 

  90. Carrasco J, Adlard P, Cotman C, Quintana A, Penkowa M, Xu F, Van Nostrand WE, Hidalgo J. Metallothionein-I and -III expression in animal models of Alzheimer disease. Neuroscience. 2006;143(4):911–22. https://doi.org/10.1016/j.neuroscience.2006.08.054.

    Article  CAS  PubMed  Google Scholar 

  91. Franceschi C, Valensin S, Lescai F, Olivieri F, Licastro F, Grimaldi LM, Monti D, De Benedictis G, Bonafe M. Neuroinflammation and the genetics of Alzheimer's disease: the search for a pro-inflammatory phenotype. Aging (Milano). 2001;13(3):163–70.

    CAS  Google Scholar 

  92. Adlard PA, West AK, Vickers JC. Increased density of metallothionein I/II-immunopositive cortical glial cells in the early stages of Alzheimer's disease. Neurobiol Dis. 1998;5(5):349–56. https://doi.org/10.1006/nbdi.1998.0203.

    Article  CAS  PubMed  Google Scholar 

  93. Cherny RA, Atwood CS, Xilinas ME, Gray DN, Jones WD, McLean CA, Barnham KJ, Volitakis I, Fraser FW, Kim Y, Huang X, Goldstein LE, Moir RD, Lim JT, Beyreuther K, Zheng H, Tanzi RE, Masters CL, Bush AI. Treatment with a copper-zinc chelator markedly and rapidly inhibits beta-amyloid accumulation in Alzheimer's disease transgenic mice. Neuron. 2001;30(3):665–76.

    Article  CAS  Google Scholar 

  94. Zhang YH, Raymick J, Sarkar S, Lahiri DK, Ray B, Holtzman D, Dumas M, Schmued LC. Efficacy and toxicity of clioquinol treatment and A-beta42 inoculation in the APP/PSI mouse model of Alzheimer's disease. Curr Alzheimer Res. 2013;10(5):494–506.

    Article  Google Scholar 

  95. Sampson EL, Jenagaratnam L, McShane R. Metal protein attenuating compounds for the treatment of Alzheimer's dementia. Cochrane Database Syst Rev. 2014;(2):CD005380. https://doi.org/10.1002/14651858.CD005380.pub5.

  96. Krishnan HS, Bernard-Gauthier V, Placzek MS, Dahl K, Narayanaswami V, Livni E, Chen Z, Yang J, Collier TL, Ran C, Hooker JM. Metal Protein-Attenuating Compound for PET Neuroimaging: Synthesis and Preclinical Evaluation of [(11)C]PBT2. 2018;15(2):695–702. https://doi.org/10.1021/acs.molpharmaceut.7b00936.

    Article  CAS  Google Scholar 

  97. Lannfelt L, Blennow K, Zetterberg H, Batsman S, Ames D, Harrison J, Masters CL, Targum S, Bush AI, Murdoch R, Wilson J, Ritchie CW, group PEs. Safety, efficacy, and biomarker findings of PBT2 in targeting Abeta as a modifying therapy for Alzheimer's disease: a phase IIa, double-blind, randomised, placebo-controlled trial. Lancet Neurol. 2008;7(9):779–86. https://doi.org/10.1016/S1474-4422(08)70167-4.

    Article  CAS  Google Scholar 

  98. Sampson EL, Jenagaratnam L, McShane R. Metal protein attenuating compounds for the treatment of Alzheimer's dementia. Cochrane Database Syst Rev. 2014;2:23. https://doi.org/10.1002/14651858.CD005380.pub5.

    Article  Google Scholar 

  99. 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(3):831–42. https://doi.org/10.1016/j.celrep.2014.06.047.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Flinn JM, Bozzelli PL, Adlard PA, Railey AM. Spatial memory deficits in a mouse model of late-onset Alzheimer's disease are caused by zinc supplementation and correlate with amyloid-beta levels. Front Aging Neurosci. 2014;6:174. https://doi.org/10.3389/fnagi.2014.00174.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Lee JY, Cole TB, Palmiter RD, Suh SW, Koh JY. Contribution by synaptic zinc to the gender-disparate plaque formation in human Swedish mutant APP transgenic mice. Proc Natl Acad Sci U S A. 2002;99(11):7705–10. https://doi.org/10.1073/pnas.092034699.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Roy S, Gumulec J, Kumar A, Raudenska M, Baig MH, Polanska H, Balvan J, Gupta M, Babula P, Odstrcilik J, Choi I, Provaznik I, Masarik M. The effect of Benzothiazolone-2 on the expression of Metallothionein-3 in modulating Alzheimer's disease. Brain Behav. 2017;7(9):e00799. https://doi.org/10.1002/brb3.799.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank Dr. Alyce Russell and Dr. Muhammad Zubair for their assistance in editing the manuscript. We apologize to those authors whose work could not be cited due to space constraints.

Funding

This work was supported by the National Natural Science Foundation of China (31571042&31671284), the Special Fund for University Teachers and the Institute for Scientific Research Cooperation at the Chinese Academy of Sciences (Y55201BY00), the Hundreds of Excellent Innovative Talents in Hebei Province Support Plan II (BR2–218), the National Postdoctoral Program for Innovative Talents (BX201600045), and the Fundamental Research Funds for the Central Universities (PA2017GDQT0018).

Author information

Authors and Affiliations

Authors

Contributions

All authors have contributed to the writing of this manuscript. Furthermore, all authors read and approved the final manuscript.

Corresponding author

Correspondence to Minglin Lang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, Y., Xiao, G., Liu, L. et al. Zinc transporters in Alzheimer’s disease. Mol Brain 12, 106 (2019). https://doi.org/10.1186/s13041-019-0528-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13041-019-0528-2

Keywords