Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-23T10:30:36.951Z Has data issue: false hasContentIssue false

Intracellular zinc in insulin secretion and action: a determinant of diabetes risk?

Published online by Cambridge University Press:  14 September 2015

Guy A. Rutter*
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
Pauline Chabosseau
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
Elisa A. Bellomo
Affiliation:
Metal Metabolism Group, Division of Diabetes and Nutritional Sciences, King's College London, School of Medicine, London SE1 9NH, UK
Wolfgang Maret
Affiliation:
Metal Metabolism Group, Division of Diabetes and Nutritional Sciences, King's College London, School of Medicine, London SE1 9NH, UK
Ryan K. Mitchell
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
David J. Hodson
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
Antonia Solomou
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
Ming Hu
Affiliation:
Section of Cell Biology and Functional Genomics, Division of Diabetes, Endocrinology and Metabolism, Imperial Centre for Translational and Experimental Medicine, Imperial College London, Du Cane Road, London W12 0NN, UK
*
*Corresponding author: G. A. Rutter, email g.rutter@imperial.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Zinc is an important micronutrient, essential in the diet to avoid a variety of conditions associated with malnutrition such as diarrhoea and alopecia. Lowered circulating levels of zinc are also found in diabetes mellitus, a condition which affects one in twelve of the adult population and whose treatments consume approximately 10 % of healthcare budgets. Zn2+ ions are essential for a huge range of cellular functions and, in the specialised pancreatic β-cell, for the storage of insulin within the secretory granule. Correspondingly, genetic variants in the SLC30A8 gene, which encodes the diabetes-associated granule-resident Zn2+ transporter ZnT8, are associated with an altered risk of type 2 diabetes. Here, we focus on (i) recent advances in measuring free zinc concentrations dynamically in subcellular compartments, and (ii) studies dissecting the role of intracellular zinc in the control of glucose homeostasis in vitro and in vivo. We discuss the effects on insulin secretion and action of deleting or over-expressing Slc30a8 highly selectively in the pancreatic β-cell, and the role of zinc in insulin signalling. While modulated by genetic variability, healthy levels of dietary zinc, and hence normal cellular zinc homeostasis, are likely to play an important role in the proper release and action of insulin to maintain glucose homeostasis and lower diabetes risk.

Type
Conference on ‘Diet, gene regulation and metabolic disease’
Copyright
Copyright © The Authors 2015 

Reflecting its importance for human health( Reference Prasad 1 ), zinc is present in all human tissues. The total amount of zinc in the human body is 2–3 g, with pancreas and bone having a particularly high content (Table 1). Total plasma zinc concentrations are 10–20 µm (Table 1) with free concentration in the blood in the nm range( Reference Magneson, Puvathingal and Ray 2 ).

Table 1. Total concentrations of zinc in various human tissues

Remarkably, almost 3000 mammalian proteins bind Zn2+ ions( Reference Andreini, Banci and Bertini 3 ), some 10 % of the proteome. Consequently, zinc plays a role in almost all aspects of cellular physiology. Aside from the structural role for tightly bound zinc in many proteins (e.g. transcription factors of the zinc finger family), these ions also modulate protein function dynamically( Reference Maret 4 ). Indeed, over 300 enzymes are dependent on zinc for catalysis( Reference Maret 4 ).

Binding of zinc to regulatory sites on proteins means that tight control of cellular zinc concentrations over a limited range is crucial for the balance between health and disease: below a certain level, zinc will be limiting, while high zinc concentrations are likely to be toxic. A plethora of proteins is consequently involved in the control of intracellular zinc in man and other mammals. Cellular homeostasis is achieved firstly by ten members of the zinc transporter (ZnT; SLC30A) family( Reference Huang and Tepaamorndech 5 ), which export Zn2+ ions from the cytosol to intracellular organelles or the extracellular space. A further fourteen members of the zinc importer (SLC39A) family( Reference Jeong and Eide 6 ) import zinc from these compartments into the cytosol (Fig. 1). Soluble metallothioneins (1–4) buffer and translocate zinc within the cytosol (see later and Fig. 4). Together, these systems achieve resting free Zn2+ concentrations in the cell cytosol in the range of 600 pm–1 nm ( Reference Vinkenborg, Nicolson and Bellomo 7 , Reference Chabosseau, Tuncay and Meur 8 ) (see next section). In some cell types, levels of free Zn2+ increase in response to stimuli( Reference Taylor, Hiscox and Nicholson 9 ), to 1 nm or above( Reference Krezel, Hao and Maret 10 , Reference Li and Maret 11 ); at the site of release from intracellular stores, free zinc concentrations might be higher. However, with some exceptions( Reference Hasan, Rink and Haase 12 , Reference Yamasaki, Sakata-Sogawa and Hasegawa 13 ), changes in cytosolic zinc are slower and smaller than those seen e.g. for calcium ions where a signalling role is very clearly defined( Reference Gilon, Chae and Rutter 14 ).

Fig. 1. (Colour online) Likely membrane topology and roles of the zinc importer (ZiP) and transporter (ZnT) families.

None of the targets of the released Zn2+ ions has so far been structurally characterised. Zinc could in theory modulate protein function in co-catalytic sites and/or by binding to and inhibiting enzymes. A well characterised zinc target is the metal responsive element transcription factor-1, which senses increased zinc concentrations and mediates zinc-dependent expression of genes such as the metallothioneins( Reference Laity and Andrews 15 ). Moreover, it has been postulated that zinc ion transients control protein–protein interactions (http://www.iospress.nl/book/zinc-in-human-health/, Chapter 4).

Imaging free Zn2+ in living cells

While total zinc levels in cells are in the mm range (Table 1)( Reference Maret 4 ), labile i.e. free concentrations are many orders of magnitude lower (pm–nm) in the range likely to regulate physiological targets. However, these concentrations are likely to differ between cell types, different intracellular organelles and in response to environmental perturbation or stimulation.

A deeper understanding of the role of zinc in cell biology and cell signalling has required the development of sensitive and non-invasive tools which provide both spatial and temporal resolution. Fluorescence microscopy is usually adopted and uses one of two probe types( Reference Dean, Qin and Palmer 16 ). First, low molecular weight compounds which display a massive change in fluorescence intensity upon zinc chelation can be loaded into the cell. The second class are genetically-encoded zinc sensors, usually expressed from an introduced cDNA and which bind Zn2+ ions through a defined metal-binding protein domain to influence the fluorescence of protein-based fluorophores to which it is fused. These zinc sensors/probes differ according to their affinity for zinc (or sensitivity, reflected by the dissociation constant K d), their selectivity for zinc against other metals ions and their dynamic range i.e. the change in fluorescence intensity triggered by zinc binding.

Chemically synthesised, low molecular weight probes can be divided in two categories: intensity-based and ratiometric (the latter will not be described here). Intensity-based probes are ‘turn-on’ fluorophores, and display a chelation-dependent increase in fluorescence intensity of up to 100-fold. Most of these probes are based on the modulation in the photon-induced electron transfer phenomenon( Reference Dean, Qin and Palmer 16 Reference de Silva, Moody and Wright 18 ). Briefly, the probes are composed of a fluorophore, a spacer domain and an electron-rich metal chelate. Photoexcitation of the fluorophore is relaxed through electron transfer with the chelate and this fluorescence quenching is supressed upon metal–ion binding. A variety of fluorescent probes with varying affinities and excitation/emission wavelengths have been described( Reference Carter, Young and Palmer 19 ). The first generation, including Zinquin( Reference Zalewski, Forbes and Betts 20 ), was derived from quinoline, a UV-excitable fluorophore. To resolve problems related to UV-excitation the next generation ZinPyr family (from ZP1 to ZP10), ZnAF, etc. was developed using fluorophores excitable with visible light. The widely-used FluoZin-3, originally based on a Ca2+ probe( Reference Gee, Zhou and Qian 21 ), displays a high affinity for zinc (K d = 15 nm). Trappable versions based on the intracellular cleavage of acetoxymethyl esters( Reference Tsien 22 ) and improved in terms of brightness and photostability, have also been developed( Reference Carter, Young and Palmer 19 ).

Compartmentalisation of these probes is not readily controlled and can vary depending on the cell type( Reference Carter, Young and Palmer 19 , Reference Qin, Miranda and Stoddard 23 , Reference Thompson, Dockery and Horobin 24 ). Biological targeting is, however, possible e.g. to mitochondria with the addition of a thiamine pyrophosphate group( Reference Chyan, Zhang and Lippard 25 ). Attachment to the cell surface to allow sampling of extracellular Zn2+ has also been achieved, and used to measure co-release of Zn2+ alongside insulin( Reference Li, Chen and Bellomo 26 , Reference Pancholi, Hodson and Kobe 27 ). The latter ‘click-SnAr-click’ strategy was also used to add targeting moieties to other intracellular localisations (e.g. mitochondria and lysosomes)( Reference Pancholi, Hodson and Kobe 27 ).

Genetically-encoded zinc sensors use Fӧrster Resonance Energy Transfer as the sensing modality and consist of a donor and acceptor fluorescent protein linked by a zinc-binding peptide sequence( Reference Hessels and Merkx 28 ). Two families have been designed by the Palmer group based on zinc fingers: Zif- and Zap-sensors. The low (μm) affinity Zif family is derived from the mammalian transcription factor Zif268, and contains either a wild type zinc fingers (ZifCY1), or a mutated (ZifCY2) domain( Reference Dittmer, Miranda and Gorski 29 , Reference Qin, Dittmer and Park 30 ). The Zap sensors, based on the Saccharomyces cerevisiae transcriptional regulator Zap1, have a very high (pm) affinity for zinc( Reference Qin, Dittmer and Park 30 ). The first member of the family ZapCY1 showed a K d of 2·5 pm, and was saturated when expressed in HeLa cells. ZapCY2 has a decreased affinity (Kd = 811 pm) and is suitable for in cellulo measurements.

The eCALWY sensors( Reference Vinkenborg, Nicolson and Bellomo 7 ) developed by Merkx and colleagues, consist of two cysteine-containing metal binding domains (ATOX1 and WD4) connected by a long, flexible glycine−serine linker and flanked by modified cerulean and citrine fluorescent proteins. Shortening of the linker length between the metal binding domains and/or mutation of one of the metal binding cysteines in the WD4 domain yielded a series of sensor variants showing affinities (pm–nm)( Reference Vinkenborg, Nicolson and Bellomo 7 ).

Taking advantage of simple fusion with a targeting sequence, several zinc sensors have been addressed to different organelles such as the mitochondria (mito-eCALWY-4, mito-ZapCY1)( Reference Chabosseau, Tuncay and Meur 8 , Reference Park, Qin and Galati 31 ), the endoplasmic reticulum (ER-eCALWY-4, ER-ZapCY1)( Reference Chabosseau, Tuncay and Meur 8 , Reference Qin, Miranda and Stoddard 23 ), Golgi apparatus (golgi-ZapCY1)( Reference Qin, Dittmer and Park 30 ) nucleus (NLS-Zaps)( Reference Miranda, Weaver and Qin 32 ) and insulin-secreting vesicles (vamp2-eCALWYs, vamp2-eZinCh1)( Reference Vinkenborg, Nicolson and Bellomo 7 ).

A summary for the results obtained with these probes is presented in Fig. 2. While similar values were returned for all probes when located in the cytosol, i.e. 0·1–1·5 nm ( Reference Vinkenborg, Nicolson and Bellomo 7 , Reference Chabosseau, Tuncay and Meur 8 , Reference Hessels and Merkx 28 , Reference Qin, Dittmer and Park 30 , Reference Bellomo, Meur and Rutter 33 ) (Fig. 2), in line with results using FluoZin-3( Reference Li and Maret 11 ), much more variation exists for mitochondria: (0·1 pm for mito-ZapCY1( Reference Park, Qin and Galati 31 ) and 300 pm for mito-eCALWY-4( Reference Chabosseau, Tuncay and Meur 8 )) and endoplasmic reticulum (about 1pm ( Reference Qin, Dittmer and Park 30 ) and above 5 nm ( Reference Chabosseau, Tuncay and Meur 8 )). The reasons for these variations are unclear and may involve differences in intracellular pH on which the probes are steeply dependent. Red-shifted variant have been created for Zap and eCALWY sensors( Reference Miranda, Weaver and Qin 32 , Reference Lindenburg, Hessels and Ebberink 34 ).

Fig. 2. (Colour online) Measurement of free Zn2+ concentrations in subcellular compartments in mammalian cells using genetically-encoded sensors. Green: ZapCY1/2( Reference Park, Qin and Galati 31 ), white: eCALWY-4( Reference Vinkenborg, Nicolson and Bellomo 7 ), yellow: eZinCh-1( Reference Vinkenborg, Nicolson and Bellomo 7 ).

Hybrid probes include both genetically-encoded and small molecular elements. These include a probe based on a carbonic anhydrase variant covalently bound to a chemical fluorophore or fused with a red fluorescent protein( Reference Bozym, Thompson and Stoddard 35 , Reference Hurst, Wang and Thompson 36 ).

The above zinc sensors and probes are thus precious tools with which to decipher the link between changes in intracellular zinc levels, diabetes risk and pathogenic mechanisms.

Zinc and diabetes

A role for zinc in diabetes aetiology has been known since 1930, when the zinc concentration was reported to be reduced by about 50 % in the pancreas of diabetic compared with non-diabetic cadavers( Reference Scott and Fisher 37 ). Epidemiological studies also suggest that whole body zinc status might be associated with diabetes( Reference el-Yazigi, Hannan and Raines 38 , Reference Garg, Gupta and Goyal 39 ). Studies on patients with type 2 diabetes (T2D) revealed that the serum concentration of zinc was decreased compared with healthy control subjects( Reference Basaki, Saeb and Nazifi 40 , Reference Jansen, Rosenkranz and Overbeck 41 ), a finding associated with increased urinary zinc loss( Reference el-Yazigi, Hannan and Raines 38 ).

Adequate levels of Zn2+ are essential not only to ensure appropriate synthesis, storage and structural stability of insulin( Reference Dunn 42 ) but also to protect against oxidative stress in T1 and T2 diabetes and their associated pathologies( Reference Cruz, de Oliveira and Marreiro 43 ). Thus, zinc is a pro-antioxidant, and a cofactor of superoxide dismutase (isoforms 1 and 3), that regulates the expression of metallothioneins and glutamate-cysteine ligase, thus affecting glutathione levels( Reference Cruz, de Oliveira and Marreiro 43 ).

The idea that zinc supplementation might improve the symptoms of T2D have been examined not only in animal models but also in diabetic patients. Studies on obese ob/ob mice showed that high zinc supplementation attenuates fasting hyperglycaemia and hyperinsulinaemia( Reference Begin-Heick, Dalpe-Scott and Rowe 44 ). Later studies provided similar results, and also revealed reduced body weight following treatment with lower zinc concentrations, in db/db mice( Reference Simon and Taylor 45 ). Moreover, recent studies on streptozotocin-induced diabetic rats revealed that zinc supplementation improves symptoms such as polydipsia and high HDL cholesterol levels( Reference Wang, Li and Fan 46 ). Amelioration of the diabetic phenotype in T2D patients has also been observed upon zinc supplementation (see Table 2. Data reviewed in( Reference Vardatsikos, Pandey and Srivastava 47 )).

Table 2. Zinc supplementation studies in Type 2 diabetic patients (modified from( Reference Vardatsikos, Pandey and Srivastava 47 ))

HbA1c, Glycosylated haemoglobin A (a time-averaged measure of blood glucose).

Zinc imaging to assess beta cell mass in vivo?

Pancreatic β-cells have an exceptionally high total Zn2+ content (about 10–20 mm), with the majority of β-cell Zn2+ being located within dense core insulin secretory granules( Reference Hutton, Penn and Peshavaria 48 ). Zinc ions are relatively less abundant in the exocrine pancreas and in other tissues( Reference Sondergaard, Stoltenberg and Doering 49 ). Although type 1 diabetes involves immune-mediated destruction of insulin-secreting pancreatic β-cells, T2D usually involves defects in both insulin release, and in hormone action( Reference Kahn 50 ). Both impaired glucose sensing by the β-cell( Reference Rutter and Parton 51 ), and a loss of overall β-cell mass( Reference Butler, Janson and Bonner-Weir 52 ), are involved in the decline in β-cell function in T2D, though the relative contributions of each are contested( Reference Rahier, Guiot and Goebbels 53 ). Classical in vivo imaging approaches such as magnetic resonance and positron emission tomography are challenging due to the relative paucity of islets within the pancreas (about 1 % of the total pancreatic volume) and the absence of suitable contrast reagents, which are thus eagerly sought. Dual-modal zinc probes, based on transition metal chelates capable of binding these ions, may therefore be of clinical value as imaging tools in the future( Reference Stasiuk, Minuzzi and Sae-Hen 54 ).

Zinc transporter 8 in insulin secretion and type 2 diabetes risk

During its biosynthesis in β-cells insulin is initially packaged as Zn2+-coordinated hexamers( Reference Arvan and Halban 55 ) before its concentration and crystallisation into dense cores within the secretory granule. The hormone is subsequently released into the circulation upon stimulation by glucose( Reference Rutter, Pullen and Hodson 56 ). The latter process involves enhanced mitochondrial metabolism( Reference Rutter, Pralong and Wollheim 57 ), closure of ATP-sensitive K+ channels( Reference Tarasov, Ravier and Semplici 58 ), Ca2+ influx( Reference Rutter, Theler and Li 59 ) and granule fusion at the plasma membrane( Reference Rutter 60 ).

ZnT8 (encoded by the SLC30A8 gene), a member of the zinc transporter family (see Fig. 1 and Introduction) is highly and almost exclusively expressed in the β- and α-cells of the endocrine pancreas, where it facilitates the import of cytosolic Zn2+ into secretory granules (Fig. 3)( Reference Chimienti, Devergnas and Favier 61 ). Autoantibodies to ZnT8 have been found in individuals with type 1 diabetes and are a relevant prognostic feature of the disease( Reference Wenzlau, Juhl and Yu 62 ). Importantly, genome-wide association studies have provided a potential link between ZnT8 activity and T2D development. Thus, Sladek et al. ( Reference Sladek, Rocheleau and Rung 63 ) found that an SNP, rs13266634, within exon 13 of the SLC30A8 gene was enriched in individuals with TD2. The variant at rs13266634 results in a missense mutation whereby an arginine residue is replaced by a tryptophan at position 325 (R325W). Risk allele carriers (R325) present with impaired insulin secretion during an intravenous glucose tolerance test( Reference Staiger, Machicao and Stefan 64 ) increased proinsulin: insulin ratio( Reference Kirchhoff, Machicao and Haupt 65 ) and lower β-cell function (as assessed through Homeostatic model assessment-B( Reference Dimas, Lagou and Barker 66 ), suggestive of impairments in both insulin secretion and processing. Four SNP located in the 3′-untranslated region of SLC30A8, two of which are in strong linkage disequilibrium with rs13266634, have also been associated with an increased risk of developing T2D. However, there are conflicting reports as to whether or not possession of these SNP is associated with changes in certain parameters of glucose homeostasis (e.g. fasting blood glucose, glucose tolerance etc.)( Reference Staiger, Machicao and Stefan 64 Reference Cauchi, Del and Choquet 68 ). Interestingly, the effects of dietary zinc supplementation to lower T2D risk are dependent on SLC30A8 genotype( Reference Kanoni, Nettleton and Hivert 69 ). More recent studies in human populations have reported twelve rare loss-of-function SNP, resulting in the production of a truncated protein that is associated with a 65 % decrease in T2D risk( Reference Flannick, Thorleifsson and Beer 70 ). This result was unexpected given the action of the common variant, which is likely to lower ZnT8 activity (see later), to increase disease risk.

Fig. 3. (Colour online) Schematic of the genome-wide association studies (GWAS)-identified zinc transporter ZnT8 and localisation of the risk associated polymorphic amino acid R/W 325.

Fig. 4. (Colour online) Possible sites of insulin action on intracellular Zn2+ homeostasis. Intracellular free zinc concentration could be impacted by: – action on zinc transporter (Zip and ZnT) on the release and uptake, respectively, of zinc ions from intracellular stores (1) or the extracellular medium (2) or by modulating the buffering of Zn2+by metallothioneins (3)( Reference Jansen, Rosenkranz and Overbeck 41 ).

Providing a link between increased T2D risk and rs13266634 inheritance has proved challenging. Unlike most genome-wide association studies identified SNP to date, rs13266634 is located in a coding region (exon) of the genome and, therefore, affects the primary sequence of ZnT8 and potentially the tertiary structure of the protein and its function (Fig. 3). However, the crystal structure of ZnT8 is yet to be elucidated and structure–function relationships, as well as the effects of SNP upon these, have been modelled on the bacterial homologue YiiP, which only shares 51·8 % sequence homology( Reference Chao and Fu 71 ). Position 325 is located towards the ‘tip’ of the ZnT8 molecule within regions involved in homodimerisation, but as both R- and W-side chains point away from the dimerization interface itself, it is unlikely that R325W substitution would have an effect on ZnT8 dimerisation or Zn2+ binding( Reference Nicolson, Bellomo and Wijesekara 72 , Reference Weijers 73 ). It is nonetheless conceivable that the positive charge of the R-side chain may hinder inter- and intra-molecular interactions, for example with presently unidentified Zn2+-binding proteins which deliver the ions to the mouth of the channel to facilitate Zn2+ transport. Providing evidence for this are zinc uptake studies( Reference Nicolson, Bellomo and Wijesekara 72 , Reference Kim, Toyofuku and Lynn 74 ), which showed that R325 is a less active Zn2+ transporter than the non-risk W325 in β-cells.

To further elucidate the role of ZnT8 in glucose homeostasis, several groups including our own have produced animal models harbouring either global ZnT8 deletion( Reference Nicolson, Bellomo and Wijesekara 72 , Reference Lemaire, Ravier and Schraenen 75 Reference Gerber, Bellomo and Hodson 77 ) or deletion restricted to the β-cell( Reference Wijesekara, Dai and Hardy 78 , Reference Tamaki, Fujitani and Hara 79 ), with some recombination in the hypothalamus( Reference Wicksteed, Brissova and Yan 80 ). Each mouse model shows variations in certain phenotypic traits (see Table 3)( Reference Rutter 81 ), which are attributed to differences in genetic background, deletion strategy and housing conditions. While the majority of animal models displayed impairments in glucose tolerance upon ZnT8 deletion, albeit with subtle age and sex-differences between the colonies, no investigators reported improvements in glucose tolerance upon ZnT8 deletion. Using electron microscopy, marked changes were seen in insulin granule morphology, with a large proportion of insulin secretory granules lacking a dense core of crystallised insulin, but containing either ‘empty’ granules or granules possessing ‘rod like’ structures. Islet Zn2+ was also decreased in both global and β-cell specific ZnT8 null animals( Reference Nicolson, Bellomo and Wijesekara 72 , Reference Wijesekara, Dai and Hardy 78 ). Although circulating insulin levels were significantly lowered in ZnT8 null mice compared with controls, glucose-stimulated insulin release was unchanged or slightly increased in islets isolated from ZnT8 null mice. Providing an elegant explanation for these apparently contradictory results, Tamaki et al.( Reference Tamaki, Fujitani and Hara 79 ) recently demonstrated a role for Zn2+ co-secreted with insulin from granules in regulating the rate of hepatic insulin clearance mediated by clathrin-dependent internalisation of the insulin receptor. As Zn2+ does not affect the uptake of C-peptide or proinsulin, this mechanism could potentially explain the impairments in both circulating insulin and the increased proinsulin:insulin ratio seen in risk allele carriers( Reference Rutter and Chimienti 82 ). Interestingly, cytosolic free Zn2+, as measured with the eCALWY4 probe, was reduced in the global knockout (KO) mouse for ZnT8( Reference Gerber, Bellomo and Hodson 77 ), alongside granule zinc concentrations as estimated by the release of zinc during exocytosis( Reference Li, Chen and Bellomo 26 ). These findings imply a more complex role for this transporter in the regulation of zinc fluxes than has previously been appreciated.

Table 3. Glycemic phenotpye of ZnT8 null mouse lines

RIP2Cre, rat insulin promoter 2-driven Cre recombinase; GSIS, glucose-stimulated insulin secretion; IGT, impaired glucose tolerance; NR, not recorded (unpublished results).

Exposing ZnT8 null animals to the diabetogenic effects of a high-fat diet led to varying results between global and β-cell specific mouse models. High-fat diet-fed β-cell specific null mice displayed similar bodyweights but displayed impaired secretion and were glucose intolerant compared with littermate controls whereas global deletion of ZnT8 resulted in increased weight gain and subsequently displayed higher levels of insulin resistance( Reference Hardy, Wijesekara and Genkin 83 ).

Notably, ZnT8 is also expressed at low but detectable levels in rat-insulin promoter-expressing neurons residing within hypothalamic appetite centres. Therefore, ectopic deletion in both global and rat-insulin promoter2Cre-driven ZnT8 mouse models may explain the phenotype observed in some of these mouse models. To overcome this issue, as well as other issues with Cre strains that express a human growth hormone minigene( Reference Brouwers, de and Osipovich 84 ), we deleted ZnT8 using a highly-specific β-cell Ins1Cre driver line, which produces no detectable recombination in the hypothalamus( Reference Thorens, Tarussio and Maestro 85 ). Confirming previous results, Ins1Cre-mediated ZnT8 deletion results in impaired glucose tolerance, abnormal insulin granule morphology and reduced cytosolic Zn2+ concentrations( Reference Mitchell, Hu and Meur 86 ). This model also shows reduced liberation of Zn2+ from isolated islets, supporting the view of Tamaki et al. that effects of Zn2+ on the liver may limit the levels of bioavailable insulin to impair glucose tolerance. Furthermore, we have recently generated a mouse line in which the protective variant (W325) of human ZnT8 is expressed selectively in the β-cell driven by an insulin promoter-controlled tetracycline responsive promoter( Reference Pullen, Sylow and Sun 87 ). In contrast to both global and β-cell-specific ZnT8 null mice, increased ZnT8 expression significantly improved glucose tolerance compared with wild type littermate controls, as expected. However, the improved glucose tolerance did not appear to be due to enhanced insulin secretion, which was reduced in islets isolated from these animals. This is likely instead to be secondary to elevations in secreted Zn2+, acting to inhibit insulin secretion through an autocrine/paracrine loop( Reference Mitchell, Hu and Meur 86 , Reference Slepchenko, James and Li 88 ). Whether the enhanced Zn2+ secretion has any effects on neighbouring α-cells, or on hepatic insulin clearance( Reference Tamaki, Fujitani and Hara 79 ), is yet to be investigated.

Genome-wide association studies and animal data have provided conflicting reports regarding the role of ZnT8 in T2D risk (see( Reference Rutter and Chimienti 82 ) for a detailed discussion). Collectively, these data would suggest that both inheritance of rs13266634 and deletion of ZnT8 in mice are detrimental in maintaining glucose homeostasis. Given the highly-restricted expression profile of ZnT8, there may be some promise in therapeutically-targeting ZnT8 in the treatment of T2D.

A role for zinc transporter 8 in glucagon secretion?

In addition to expression in β-cells, there is evidence that ZnT8 is also present in α-cells, at least in mouse and human pancreata. Its expression in these cells was confirmed by immunocytochemistry in pancreatic slices and dissociated islets and by gene expression analysis on a fluorescence-activated cell sorting purified mouse population( Reference Nicolson, Bellomo and Wijesekara 72 ). Similarly to the β-cell, ZnT8 appears to be the most highly expressed member of the ZnT family in the α-cell. In the porcine pancreas, however, ZnT8 is exclusively expressed in the β-cell( Reference Schweiger, Steffl and Amselgruber 89 ) implying clear species differences and suggesting that in pig islets zinc homeostasis is regulated differently from human and rodent islets. Knockdown of ZnT8 in glucagonoma-derived αTC1·9 cells resulted in an increase of glucagon mRNA, and decreased regulated glucagon secretion by 70 %. Overexpression of the ZnT8 R325 or W325 variants in these cells reciprocally led to reduced glucagon content and 50 % lower glucagon secretion( Reference Souza, Qui and Inouye 90 ).

As a result of its co-secretion with insulin, zinc can reach high local concentrations within the islet, with the potential of acting as a mediator of paracrine signalling. Thus, in the perfused rat pancreas, the zinc chelator Ca2+-EDTA led to a stimulation of secretion when using the mitochondrial substrate monoethyl-succinate as a secretagogue, an agent which normally acts as an inhibitor of glucagon release( Reference Ishihara, Maechler and Gjinovci 91 ). Correspondingly, zinc diminished pyruvate- or glucose-induced glucagon secretion from isolated rat α-cells through the reversible activation of KATP channels and a subsequent decrease in electrical activity( Reference Franklin, Gromada and Gjinovci 92 ). Likewise, studies in the mouse demonstrated an inhibitory effect of Zn2+ on glucagon secretion( Reference Gyulkhandanyan, Lu and Lee 93 ). However, in the latter species, this effect of zinc could not be attributed to KATP channel activation but to uptake of the ion by calcium channels and an alteration in redox state( Reference Gyulkhandanyan, Lu and Lee 93 ). Of note, fasting plasma glucagon levels were normal in global ZnT8 null mice and glucagon secretion from isolated islets was not higher in KO mice than controls. However, exogenous zinc did lead to a reduction of glucagon secretion under stimulatory conditions( Reference Hardy, Serino and Wijesekara 94 ). Assuming that granular zinc was almost entirely depleted in the KO mice, these data suggest that zinc secreted from β-cells is not responsible for the inhibition of glucagon secretion at high glucose.

Although there have been extensive studies on the role of ZnT8 in the β-cell (see preceding section and references therein), its function in the α-cell has not been investigated in detail. One study looked briefly at an α-cell-specific ZnT8 KO and did not detect any differences in fasting plasma glucagon levels or glucose homeostasis compared with control mice( Reference Wijesekara, Dai and Hardy 78 ). Further examination of islets from these mice is necessary as well as investigating changes at the single cell level.

In line with our earlier findings( Reference Wijesekara, Dai and Hardy 78 ), we have more recently observed no effect of ZnT8 deletion selectively in α-cells on glucose homeostasis or fasting glucagon( Reference Solomou, Meur and Bellomo 95 ). However, we observed in these more recent studies that KO mice displayed enhanced responses to hypoglycaemia in vivo and increased glucagon secretion from isolated islets, implying cell-autonomous roles for ZnT8 in the α-cell.

Zinc and insulin action

Insulin binding to the α-subunit of the insulin receptor leads to enhanced intrinsic protein tyrosine kinase activity of the β-subunit and phosphorylation of the receptor on multiple tyrosine residues( Reference Lee and Pilch 96 ). The activated insulin receptor then phosphorylates several scaffolding proteins, including the insulin receptor substrates, which subsequently bind and activate other signalling proteins to trigger two main different pathways (Fig. 5): the activation of the serine/threonine kinases Ras and Raf by the SOS/Grb2/SHC complex leading to extracellularly-regulated kinases1/2 activation and cellular proliferation, and activation of phosphatidyl inositol 3′-kinase( Reference Kanzaki 97 ). The latter triggers the translocation of phoshoinositide-dependent kinase-1 to the plasma membrane together with protein kinase B and activates downstream protein kinases such as p70 ribosomal S6 kinase. The latter stimulate glucose transport, glycogen synthesis, lipogenesis and other processes (Fig. 5).

Fig. 5. (Colour online) Possible sites of action of Zn2+ on insulin signalling. PTB1B, protein tyrosine phosphatase 1B; SHC, Src-homology-2-containing; SOS, Son-of-sevenless; MEK, MAP/ERK kinase; MAP, mitogen-activated protein kinase; IRS, insulin receptor substrate; P1-3K, phosphatidylinositol 3’ kinase; PTEN, phosphatase and tensin homologue; PDK, phosphoinositide-dependent kinase; aPKC, atypical protein kinase C; GSK, glycogen synthase kinase; PPI, protein phosphatase inhibitor-1; GLUT4, glucose transporter family member 4.

An action of zinc on insulin-target tissue was described for the first time by Coulston and Dandona, when they observed that treatment of rat adipocytes with high zinc concentrations led to an increased rate of lipogenesis( Reference Coulston and Dandona 98 ). At the cellular level, zinc was found to increase insulin receptor substrates-1 tyrosine phosphorylation in the presence or absence of insulin in skeletal muscle cells( Reference Miranda and Dey 99 ). Moreover, zinc was able to activate protein kinase B in several studies using 3T3-L1 and rat adipocytes: zinc treatment induced not only phosphorylation of the insulin receptor, but also of protein kinase B. While the former is possibly dependent on the inhibition by zinc of protein tyrosine phosphatase 1B( Reference Haase and Maret 100 ), the latter is a consequence of phosphatidyl inositol 3′-kinase inhibition( Reference Tang and Shay 101 ). Phosphatase and tensin homologue is responsible for the negative regulation of phosphatidyl inositol 3′-kinase /protein kinase B activity and appears to be degraded in response to zinc( Reference Wu, Wang and Zhang 102 ) at physiological (about 600 pm) concentrations( Reference Plum, Brieger and Engelhardt 103 ).

As mentioned earlier, one of the most studied targets of the insulin-mimetic effects of zinc is protein tyrosine phosphatase 1B and a recent study has elucidated the inhibition constant and the mechanisms of zinc binding to protein tyrosine phosphatase 1B ( Reference Haase and Maret 100 , Reference Bellomo, Massarotti and Hogstrand 104 ). Zinc also inhibits protein tyrosine phosphatase 1B with an apparent inhibition constant as low as 5 nm. By acting through these two targets, an increase in intracellular zinc signal can be predicted to enhance insulin signalling downstream of the insulin receptor. Zinc also mimics insulin action by triggering nuclear exclusion of the transcription factor FOXO1, and induces glycogen synthesis by lowering the phosphorylation of glycogen synthase kinase-3( Reference Barthel, Ostrakhovitch and Walter 105 ).

The effect of zinc on insulin signalling described above may also explain, at least in part, the impact of the metal ion in the pathogenesis of T2D. Although in vitro analyses showed that insulin leads to an increase in intracellular zinc( Reference Jansen, Rosenkranz and Overbeck 41 ), the exact physiological mechanisms responsible are not entirely understood (Fig. 2). While the free concentration of Zn2+ within the endoplasmic reticulum is contested (see earlier) it has been suggested that zinc importer-7-mediated release from this compartment may be involved in the actions of epidermal growth factor, which like insulin acts through a receptor tyrosine kinase( Reference Taylor, Hiscox and Nicholson 9 ). Moreover, ablation of zinc importer-7 in skeletal muscle cells displayed a reduction in GLUT4 protein expression and insulin-stimulated glycogen synthesis( Reference Myers, Nield and Chew 106 ). As discussed earlier, zinc may also inhibit insulin clearance by the liver, leading to elevated circulating levels of the hormone( Reference Tamaki, Fujitani and Hara 79 ).

Finally, zinc may also contribute to the anti-oxidant actions of insulin, serving (i) as a cofactor of superoxide dismutase, (ii) to enhance the expression of metallothioneins and (iii) to stimulate glutamate-cysteine ligase and glutathione synthesis. Interestingly, supplementing diabetic rats with zinc increased superoxide dismutase activity( Reference Zhu, Nie and Li 107 ). Decreased lipid peroxidation, concomitant with an increase in glutathione concentration, was also found in zinc treated diabetic rats( Reference Karatug, Kaptan and Bolkent 108 ). Potential sites of action of Zn2+ on insulin signalling are shown in Fig. 5.

Conclusions

Recent advances, stretching from human genetics through mouse models to the creation of new imaging modalities, have provided unexpected insights on the role of zinc in the release and actions of insulin. Importantly these new tools provide compelling evidence to indicate that these ions play an under-appreciated role to support the actions of the hormone on target tissues and suggest that Zn2+ should be considered as both an important extra- and intra-cellular signalling species. Nonetheless, important controversies, including the role of ZnT8 in the glucose homeostasis and diabetes risk in man, remain to be resolved( Reference Rutter and Chimienti 82 ).

Acknowledgements

We thank Maarten Merkx (Eindhoven University), Dax Fu (Johns Hopkins University), Christer Hogstrand (King's College London), Michael Wheeler (Toronto University), Fabrice Chimienti (AstraZeneca), Nicolas Long (Imperial College), Michael Watkinson (Queen Mary, University of London), and Graeme Stasiuk (University of Hull) for useful discussion.

Financial Support

G. A. R. was supported by a Wellcome Trust Senior Investigator Award (WT098424AIA), MRC Programme Grant (MR/J0003042/1) and a Royal Society Wolfson Research Merit Award. The work leading to this publication has received support from the Innovative Medicines Initiative Joint Undertaking under grant agreement No. 155005 (IMIDIA), resources of which are composed of financial contribution from the European Union's Seventh Framework Programme (FP7/2007–2013) and EFPIA companies’ in kind contribution (G. A. R.). D. J. H. thanks Diabetes UK for R.D. Lawrence Fellowship (12/0004431) and W. M. the Biotechnology and Biological Sciences Research Council for project grant (BB/K001442/1).

Conflicts of Interest

None.

Authorship

G. A. R. designed and coordinated the writing and assembled the final manuscript. Chief contributors to the individual sections were: G. A. R., Abstract, Introduction, Conclusions; P. C., Zn2+ imaging; E. A. B., Introduction and insulin action; W. M., Introduction; R. K. M., D. J. H., A. S. and M. H., ZnT8.

References

1. Prasad, AS (2013) Discovery of human zinc deficiency: its impact on human health and disease. Adv Nutr 4, 176190.Google Scholar
2. Magneson, GR, Puvathingal, JM & Ray, WJ Jr (1987) The concentrations of free Mg2+ and free Zn2+ in equine blood plasma. J Biol Chem 262, 1114011148.CrossRefGoogle ScholarPubMed
3. Andreini, C, Banci, LF, Bertini, IF et al. (2006) Counting the zinc-proteins encoded in the human genome. J Proteome Res 5, 196201.CrossRefGoogle ScholarPubMed
4. Maret, W (2013) Zinc biochemistry: from a single zinc enzyme to a key element of life. Adv Nutr 4, 8291.CrossRefGoogle ScholarPubMed
5. Huang, L & Tepaamorndech, S (2013) The SLC30 family of zinc transporters - a review of current understanding of their biological and pathophysiological roles. Mol Aspects Med 34, 548560.CrossRefGoogle ScholarPubMed
6. Jeong, J & Eide, DJ (2013) The SLC39 family of zinc transporters. Mol Aspects Med 34, 612619.CrossRefGoogle ScholarPubMed
7. Vinkenborg, JL, Nicolson, TJ, Bellomo, EA et al. (2009) Genetically encoded FRET sensors to monitor intracellular Zn2+ homeostasis. Nat Methods 6, 737740.CrossRefGoogle ScholarPubMed
8. Chabosseau, P, Tuncay, E, Meur, G et al. (2014) Mitochondrial and ER-targeted eCALWY probes reveal high levels of free Zn2+ . ACS Chem Biol 9, 21112120.CrossRefGoogle ScholarPubMed
9. Taylor, KM, Hiscox, S, Nicholson, RI et al. (2012) Protein kinase CK2 triggers cytosolic zinc signaling pathways by phosphorylation of zinc channel ZIP7. Sci Signal 5, ra11.Google Scholar
10. Krezel, A, Hao, Q & Maret, W (2007) The zinc/thiolate redox biochemistry of metallothionein and the control of zinc ion fluctuations in cell signaling. Arch Biochem Biophys 463, 188200.Google Scholar
11. Li, Y & Maret, W (2009) Transient fluctuations of intracellular zinc ions in cell proliferation. Exp Cell Res 315, 24632470.Google Scholar
12. Hasan, R, Rink, L & Haase, H (2013) Zinc signals in neutrophil granulocytes are required for the formation of neutrophil extracellular traps. Innate Immun 19, 253264.Google Scholar
13. Yamasaki, S, Sakata-Sogawa, K, Hasegawa, A et al. (2007) Zinc is a novel intracellular second messenger. J Cell Biol 177, 637645.CrossRefGoogle ScholarPubMed
14. Gilon, P, Chae, HY, Rutter, GA et al. (2014) Calcium signaling in pancreatic beta-cells in health and in Type 2 diabetes. Cell Calcium 56, 340361.Google Scholar
15. Laity, JH & Andrews, GK (2007) Understanding the mechanisms of zinc-sensing by metal-response element binding transcription factor-1 (MTF-1). Arch Biochem Biophys 463, 201210.Google Scholar
16. Dean, KM, Qin, Y & Palmer, AE (2012) Visualizing metal ions in cells: an overview of analytical techniques, approaches, and probes. Biochim Biophys Acta 1823, 14061415.CrossRefGoogle ScholarPubMed
17. de Silva, AP, Fox, DB, Moody, TS et al. (2001) The development of molecular fluorescent switches. Trends Biotechnol 19, 2934.Google Scholar
18. de Silva, AP, Moody, TS & Wright, GD (2009) Fluorescent PET (photoinduced electron transfer) sensors as potent analytical tools. The Analyst 134, 23852393.CrossRefGoogle ScholarPubMed
19. Carter, KP, Young, AM & Palmer, AE (2014) Fluorescent sensors for measuring metal ions in living systems. Chem Rev 114, 45644601.CrossRefGoogle ScholarPubMed
20. Zalewski, PD, Forbes, IJ & Betts, WH (1993) Correlation of apoptosis with change in intracellular labile Zn(II) using zinquin [(2-methyl-8-p-toluenesulphonamido-6-quinolyloxy)acetic acid], a new specific fluorescent probe for Zn(II). The Biochem J 296, 403408.CrossRefGoogle Scholar
21. Gee, KR, Zhou, ZL, Qian, WJ et al. (2002) Detection and imaging of zinc secretion from pancreatic beta-cells using a new fluorescent zinc indicator. J Am Chem Soc 124, 776778.CrossRefGoogle ScholarPubMed
22. Tsien, RY (1981) A non-disruptive technique for loading calcium buffers and indicators into cells. Nature 290, 527528.CrossRefGoogle ScholarPubMed
23. Qin, Y, Miranda, JG, Stoddard, CI et al. (2013) Direct comparison of a genetically encoded sensor and small molecule indicator: implications for quantification of cytosolic Zn(2+). ACS Chem Biol 8, 23662371.Google Scholar
24. Thompson, K, Dockery, P & Horobin, RW (2012) Predicting and avoiding subcellular compartmentalization artifacts arising from acetoxymethyl ester calcium imaging probes. The case of fluo-3 AM and a general account of the phenomenon including a problem avoidance chart. Biotech Histochem: Official Pub Biological Stain Commission 87, 468483.CrossRefGoogle ScholarPubMed
25. Chyan, W, Zhang, DY, Lippard, SJ et al. (2014) Reaction-based fluorescent sensor for investigating mobile Zn2+ in mitochondria of healthy versus cancerous prostate cells. Proc Natl Acad Sci 111, 143148.CrossRefGoogle ScholarPubMed
26. Li, D, Chen, S, Bellomo, EA et al. (2011) Imaging dynamic insulin release using a fluorescent zinc indicator for monitoring induced exocytotic release (ZIMIR). Proc Natl Acad Sci USA 108, 2106321068.CrossRefGoogle ScholarPubMed
27. Pancholi, J, Hodson, DJ, Kobe, K et al. (2014) Biologically targeted probes for zinc: a diversity-oriented modular “Click-SNAR” Approach 5, 35283535.Google Scholar
28. Hessels, AM & Merkx, M (2015) Genetically-encoded FRET-based sensors for monitoring Zn(2+) in living cells. Metallomics: Integr Biometal Sci 7, 258266.Google Scholar
29. Dittmer, PJ, Miranda, JG, Gorski, JA et al. (2009) Genetically encoded sensors to elucidate spatial distribution of cellular zinc. J Biol Chem 284, 1628916297.CrossRefGoogle ScholarPubMed
30. Qin, Y, Dittmer, PJ, Park, JG et al. (2011) Measuring steady-state and dynamic endoplasmic reticulum and Golgi Zn2+ with genetically encoded sensors. Proc Natl Acad Sci USA 108, 73517356.CrossRefGoogle ScholarPubMed
31. Park, JG, Qin, Y, Galati, DF et al. (2012) New sensors for quantitative measurement of mitochondrial Zn(2+). ACS Chem Biol 7, 16361640.Google Scholar
32. Miranda, JG, Weaver, AL, Qin, Y et al. (2012) New alternately colored FRET sensors for simultaneous monitoring of Zn(2)(+) in multiple cellular locations. PLoS ONE 7, e49371.Google Scholar
33. Bellomo, EA, Meur, G & Rutter, GA (2011) Glucose regulates free cytosolic Zn2+ concentration, Slc39 (ZiP) and metallothionein gene expression in primary pancreatic islet {beta}-cells. J Biol Chem 286, 2577825789.CrossRefGoogle ScholarPubMed
34. Lindenburg, LH, Hessels, AM, Ebberink, EH et al. (2013) Robust red FRET sensors using self-associating fluorescent domains. ACS Chem Biol 8, 21332139.Google Scholar
35. Bozym, RA, Thompson, RB, Stoddard, AK et al. (2006) Measuring picomolar intracellular exchangeable zinc in PC-12 cells using a ratiometric fluorescence biosensor. ACS Chem Biol 1, 103111.Google Scholar
36. Hurst, TK, Wang, D, Thompson, RB et al. (2010) Carbonic anhydrase II-based metal ion sensing: advances and new perspectives. Biochimica et Biophysica acta 1804, 393403.Google Scholar
37. Scott, DA & Fisher, AM (1938) The Insulin and the zinc content of normal and diabetic pancreas. J Clin Invest 17, 725728.Google Scholar
38. el-Yazigi, A, Hannan, N & Raines, DA (1993) Effect of diabetic state and related disorders on the urinary excretion of magnesium and zinc in patients. Diabetes Res 22, 6775.Google ScholarPubMed
39. Garg, VK, Gupta, R & Goyal, RK (1994) Hypozincemia in diabetes mellitus. J Assoc Physicians India 42, 720721.Google ScholarPubMed
40. Basaki, M, Saeb, M, Nazifi, S et al. (2012) Zinc, copper, iron, and chromium concentrations in young patients with type 2 diabetes mellitus. Biol Trace Elem Res 148, 161164.Google Scholar
41. Jansen, J, Rosenkranz, E, Overbeck, S et al. (2012) Disturbed zinc homeostasis in diabetic patients by in vitro and in vivo analysis of insulinomimetic activity of zinc. J Nutr Biochem 23, 14581466.CrossRefGoogle ScholarPubMed
42. Dunn, MF (2005) Zinc-ligand interactions modulate assembly and stability of the insulin hexamer – a review. Biometals 18, 295303.Google Scholar
43. Cruz, KJ, de Oliveira, AR & Marreiro, DN (2015) Antioxidant role of zinc in diabetes mellitus. World J Diabetes 6, 333337.Google Scholar
44. Begin-Heick, N, Dalpe-Scott, M, Rowe, J et al. (1985) Zinc supplementation attenuates insulin secretory activity in pancreatic islets of the ob/ob mouse. Diabetes 34, 179184.Google Scholar
45. Simon, SF & Taylor, CG (2001) Dietary zinc supplementation attenuates hyperglycemia in db/db mice. Exp Biol Med (Maywood) 226, 4351.Google Scholar
46. Wang, X, Li, H, Fan, Z et al. (2012) Effect of zinc supplementation on type 2 diabetes parameters and liver metallothionein expressions in Wistar rats. J Physiol Biochem 68, 563572.CrossRefGoogle ScholarPubMed
47. Vardatsikos, G, Pandey, NR & Srivastava, AK (2013) Insulino-mimetic and anti-diabetic effects of zinc. J Inorg Biochem 120, 817. Epub@2012 Dec 3., 8–17.Google Scholar
48. Hutton, JC, Penn, EJ & Peshavaria, M (1983) Low-molecular-weight constituents of isolated insulin-secretory vesicles. Bivalent cations, adenine nucleotides and inorganic phosphate. Biochem J 210, 297305.Google Scholar
49. Sondergaard, LG, Stoltenberg, M, Doering, P et al. (2006) Zinc ions in the endocrine and exocrine pancreas of zinc deficient rats. Histol Histopathol 21, 619625.Google ScholarPubMed
50. Kahn, SE (2003) The relative contributions of insulin resistance and beta-cell dysfunction to the pathophysiology of Type 2 diabetes. Diabetologia 46, 319.Google Scholar
51. Rutter, GA & Parton, LE (2008) The beta-cell in type 2 diabetes and in obesity. Front Horm Res 36, 118134.CrossRefGoogle ScholarPubMed
52. Butler, AE, Janson, J, Bonner-Weir, S et al. (2003) Beta-cell deficit and increased beta-cell apoptosis in humans with type 2 diabetes. Diabetes 52, 102110.Google Scholar
53. Rahier, J, Guiot, Y, Goebbels, RM et al. (2008) Pancreatic beta-cell mass in European subjects with type 2 diabetes. Diabetes Obes Metab 10, 3242.Google Scholar
54. Stasiuk, GJ, Minuzzi, F, Sae-Hen, M et al. (2015) Dual-modal MR/fluorescent zinc probes towards pancreatic beta cell mass imaging. Chemistry 21, 50235033.Google Scholar
55. Arvan, P & Halban, PA (2004) Sorting ourselves out: seeking consensus on trafficking in the beta-cell. Traffic 5, 5361.Google Scholar
56. Rutter, GA, Pullen, TJ, Hodson, DJ et al. (2015) Pancreatic beta cell identity, glucose sensing and the control of insulin secretion. Biochem J 466, 202218.Google Scholar
57. Rutter, GA, Pralong, W-F & Wollheim, CB (1992) Regulation of mitochondrial glycerol phosphate dehydrogenase activity by Ca2+ within electropermeabilized insulin-secreting cells (INS1). Biochimica et Biophysica Acta – Bioenergetics 1175, 107113.Google Scholar
58. Tarasov, AI, Ravier, MA, Semplici, F et al. (2012) The mitochondrial Ca2+ uniporter MCU is essential for glucose-induced ATP increases in pancreatic â-cells. PLoS ONE 7, e39722.CrossRefGoogle Scholar
59. Rutter, GA, Theler, J-M, Li, G et al. (1994) Ca2+ stores in insulin-secreting cells: lack of effect of cADP ribose. Cell Calcium 16, 7180.Google Scholar
60. Rutter, GA (2004) Visualising insulin secretion. The minkowski lecture 2004. Diabetologia 47, 18611872.Google Scholar
61. Chimienti, F, Devergnas, S, Favier, A et al. (2004) Identification and cloning of a beta-cell-specific zinc transporter, ZnT-8, localized into insulin secretory granules. Diabetes 53, 23302337.Google Scholar
62. Wenzlau, JM, Juhl, K, Yu, L et al. (2007) The cation efflux transporter ZnT8 (Slc30A8) is a major autoantigen in human type 1 diabetes. Proc Natl Acad Sci USA 104, 1704017045.Google Scholar
63. Sladek, R, Rocheleau, G, Rung, J et al. (2007) A genome-wide association study identifies novel risk loci for type 2 diabetes. Nature 445, 881885.Google Scholar
64. Staiger, H, Machicao, F, Stefan, N et al. (2007) Polymorphisms within novel risk loci for type 2 diabetes determine beta-cell function. PLoS ONE 2, e832.Google Scholar
65. Kirchhoff, K, Machicao, F, Haupt, A et al. (2008) Polymorphisms in the TCF7L2, CDKAL1 and SLC30A8 genes are associated with impaired proinsulin conversion. Diabetologia 51, 597601.CrossRefGoogle ScholarPubMed
66. Dimas, AS, Lagou, V, Barker, A et al. (2014) Impact of type 2 diabetes susceptibility variants on quantitative glycemic traits reveals mechanistic heterogeneity. Diabetes 63, 21582171.CrossRefGoogle ScholarPubMed
67. Ruchat, SM, Elks, CE, Loos, RJ et al. (2009) Association between insulin secretion, insulin sensitivity and type 2 diabetes susceptibility variants identified in genome-wide association studies. Acta Diabetol 46, 217226.CrossRefGoogle ScholarPubMed
68. Cauchi, S, Del, GS, Choquet, H et al. (2010) Meta-analysis and functional effects of the SLC30A8 rs13266634 polymorphism on isolated human pancreatic islets. Mol Genet Metab 100, 7782.CrossRefGoogle ScholarPubMed
69. Kanoni, S, Nettleton, JA, Hivert, MF et al. (2011) Total zinc intake may modify the glucose-raising effect of a zinc transporter (SLC30A8) variant: a 14-Cohort meta-analysis. Diabetes 60, 24072416.Google Scholar
70. Flannick, J, Thorleifsson, G, Beer, NL et al. (2014) Loss-of-function mutations in SLC30A8 protect against type 2 diabetes. Nat Genet 46, 357363.Google Scholar
71. Chao, Y & Fu, D (2004) Thermodynamic studies of the mechanism of metal binding to the Escherichia coli zinc transporter YiiP. J Biol Chem 279, 1717317180.Google Scholar
72. Nicolson, TJ, Bellomo, EA, Wijesekara, N et al. (2009) Insulin storage and glucose homeostasis in mice null for the granule zinc transporter ZnT8 and studies of the type 2 diabetes-associated variants. Diabetes 58, 20702083.CrossRefGoogle ScholarPubMed
73. Weijers, RN (2010) Three-dimensional structure of beta-cell-specific zinc transporter, ZnT-8, predicted from the type 2 diabetes-associated gene variant SLC30A8 R325W. Diabetol Metab Syndr 2, 33.CrossRefGoogle ScholarPubMed
74. Kim, H, Toyofuku, Y, Lynn, FC et al. (2010) Serotonin regulates pancreatic beta cell mass during pregnancy. Nat Med 16, 804808.Google Scholar
75. Lemaire, K, Ravier, MA, Schraenen, A et al. (2009) Insulin crystallization depends on zinc transporter ZnT8 expression, but is not required for normal glucose homeostasis in mice. Proc Natl Acad Sci USA 106, 1487214877.Google Scholar
76. Pound, LD, Sarkar, SA, Benninger, RK et al. (2009) Deletion of the mouse Slc30a8 gene encoding zinc transporter-8 results in impaired insulin secretion. Biochem J 421, 371376.Google Scholar
77. Gerber, PA, Bellomo, EA, Hodson, DJ et al. (2014) Hypoxia lowers SLC30A8/ZnT8 expression and free cytosolic Zn2+ in pancreatic beta cells. Diabetologia 57, 16351644.CrossRefGoogle ScholarPubMed
78. Wijesekara, N, Dai, FF, Hardy, AB et al. (2010) Beta cell specific ZnT8 deletion in mice causes marked defects in insulin processing, crystallisation and secretion. Diabetologia 53, 16561668.CrossRefGoogle ScholarPubMed
79. Tamaki, M, Fujitani, Y, Hara, A et al. (2013) The diabetes-susceptible gene SLC30A8/ZnT8 regulates hepatic insulin clearance. J Clin Invest 123, 45134524.Google Scholar
80. Wicksteed, B, Brissova, M, Yan, W et al. (2010) Conditional gene targeting in mouse pancreatic {beta}-cells: analysis of ectopic cre transgene expression in the brain. Diabetes 59, 30903098.Google Scholar
81. Rutter, GA (2010) Think zinc: new roles for zinc in the control of insulin secretion. Islets 2, 12.CrossRefGoogle ScholarPubMed
82. Rutter, GA & Chimienti, F (2015) SLC30A8 mutations in type 2 diabetes. Diabetologia 58, 3136.Google Scholar
83. Hardy, AB, Wijesekara, N, Genkin, I et al. (2012) Effects of high-fat diet feeding on Znt8-null mice: differences between beta-cell and global knockout of Znt8. Am J Physiol Endocrinol Metab 302, E1084E1096.Google Scholar
84. Brouwers, B, de, FG, Osipovich, AB et al. (2014) Impaired islet function in commonly used transgenic mouse lines due to human growth hormone minigene expression. Cell Metab 20, 979990.Google Scholar
85. Thorens, B, Tarussio, D, Maestro, MA et al. (2015) Ins1 knock-in mice for beta cell-specific gene recombination. Diabetologia 58, 558656.CrossRefGoogle ScholarPubMed
86. Mitchell, RK, Hu, M, Meur, G et al. (2015) Reciprocal changes in glucose tolerance after pancreatic beta cell-selective deletion or over-expression of Slc30a8/ZnT8 in mice Diabetologia (In the Press).Google Scholar
87. Pullen, TJ, Sylow, L, Sun, G et al. (2012) Over-expression of Monocarboxylate transporter-1 (Slc16a1) in the pancreatic â-cell leads to relative hyperinsulinism during exercise. Diabetes 61, 17191725.CrossRefGoogle Scholar
88. Slepchenko, KG, James, CB & Li, YV (2013) Inhibitory effect of zinc on glucose-stimulated zinc/insulin secretion in an insulin-secreting beta-cell line. Exp Physiol 98, 13011311.CrossRefGoogle Scholar
89. Schweiger, M, Steffl, M & Amselgruber, WM (2013) The zinc transporter ZnT8 (slc30A8) is expressed exclusively in beta cells in porcine islets. Histochem Cell Biol 140, 677684.Google Scholar
90. Souza, SC, Qui, L, Inouye, K et al. (2008) Zinc transporter ZnT-8 regulates insulin and glucagon secretion in Min6 and alphaTC1-9 pancreatic cell lines. Diabetologia 51, S206.Google Scholar
91. Ishihara, H, Maechler, P, Gjinovci, A et al. (2003) Islet beta-cell secretion determines glucagon release from neighbouring alpha-cells. Nat Cell Biol 5, 330335.CrossRefGoogle ScholarPubMed
92. Franklin, I, Gromada, J, Gjinovci, A et al. (2005) Beta-cell secretory products activate alpha-cell ATP-dependent potassium channels to inhibit glucagon release. Diabetes 54, 18081815.Google Scholar
93. Gyulkhandanyan, AV, Lu, H, Lee, SC et al. (2008) Investigation of transport mechanisms and regulation of intracellular Zn2+ in pancreatic alpha-cells. J Biol Chem 283, 1018410197.Google Scholar
94. Hardy, AB, Serino, AS, Wijesekara, N et al. (2011) Regulation of glucagon secretion by zinc: lessons from the beta cell-specific Znt8 knockout mouse model. Diabetes Obes Metab 13 (Suppl. 1), 112117.Google Scholar
95. Solomou, A, Meur, G, Bellomo, EA et al. (2015) The zinc transporter Slc30a8/ZnT8 is required in a subpopulation of pancreatic á cells for hypoglycemia-induced glucagon secretion. J Biol Chem (Epublication ahead of print version).Google Scholar
96. Lee, J & Pilch, PF (1994) The insulin receptor: structure, function, and signaling. Am J Physiol 266, C319C334.Google Scholar
97. Kanzaki, M (2006) Insulin receptor signals regulating GLUT4 translocation and actin dynamics. Endocr J 53, 267293.CrossRefGoogle ScholarPubMed
98. Coulston, L & Dandona, P (1980) Insulin-like effect of zinc on adipocytes. Diabetes 29, 665667.Google Scholar
99. Miranda, ER & Dey, CS (2004) Effect of chromium and zinc on insulin signaling in skeletal muscle cells. Biol Trace Elem Res 101, 1936.CrossRefGoogle ScholarPubMed
100. Haase, H & Maret, W (2003) Intracellular zinc fluctuations modulate protein tyrosine phosphatase activity in insulin/insulin-like growth factor-1 signaling. Exp Cell Res 291, 289298.CrossRefGoogle ScholarPubMed
101. Tang, X & Shay, NF (2001) Zinc has an insulin-like effect on glucose transport mediated by phosphoinositol-3-kinase and Akt in 3T3-L1 fibroblasts and adipocytes. J Nutr 131, 14141420.Google Scholar
102. Wu, W, Wang, X, Zhang, W et al. (2003) Zinc-induced PTEN protein degradation through the proteasome pathway in human airway epithelial cells. J Biol Chem 278, 2825828263.Google Scholar
103. Plum, LM, Brieger, A, Engelhardt, G et al. (2014) PTEN-inhibition by zinc ions augments interleukin-2-mediated Akt phosphorylation. Metallomics 6, 12771287.Google Scholar
104. Bellomo, E, Massarotti, A, Hogstrand, C et al. (2014) Zinc ions modulate protein tyrosine phosphatase 1B activity. Metallomics 6, 12291239.CrossRefGoogle ScholarPubMed
105. Barthel, A, Ostrakhovitch, EA, Walter, PL et al. (2007) Stimulation of phosphoinositide 3-kinase/Akt signaling by copper and zinc ions: mechanisms and consequences. Arch Biochem Biophys 463, 175182.Google Scholar
106. Myers, SA, Nield, A, Chew, GS et al. (2013) The zinc transporter, Slc39a7 (Zip7) is implicated in glycaemic control in skeletal muscle cells. PLoS ONE 8, e79316.Google Scholar
107. Zhu, K, Nie, S, Li, C et al. (2013) Antidiabetic and pancreas-protective effects of zinc threoninate chelate in diabetic rats may be associated with its antioxidative stress ability. Biol Trace Elem Res 153, 291298.CrossRefGoogle ScholarPubMed
108. Karatug, A, Kaptan, E, Bolkent, S et al. (2013) Alterations in kidney tissue following zinc supplementation to STZ-induced diabetic rats. J Trace Elem Med Biol 27, 5257.Google Scholar
109. Pound, LD, Sarkar, SA, Ustione, A et al. (2012) The physiological effects of deleting the mouse slc30a8 gene encoding zinc transporter-8 are influenced by gender and genetic background. PLoS ONE 7, e40972.Google Scholar
Figure 0

Table 1. Total concentrations of zinc in various human tissues

Figure 1

Fig. 1. (Colour online) Likely membrane topology and roles of the zinc importer (ZiP) and transporter (ZnT) families.

Figure 2

Fig. 2. (Colour online) Measurement of free Zn2+ concentrations in subcellular compartments in mammalian cells using genetically-encoded sensors. Green: ZapCY1/2(31), white: eCALWY-4(7), yellow: eZinCh-1(7).

Figure 3

Table 2. Zinc supplementation studies in Type 2 diabetic patients (modified from(47))

Figure 4

Fig. 3. (Colour online) Schematic of the genome-wide association studies (GWAS)-identified zinc transporter ZnT8 and localisation of the risk associated polymorphic amino acid R/W 325.

Figure 5

Fig. 4. (Colour online) Possible sites of insulin action on intracellular Zn2+ homeostasis. Intracellular free zinc concentration could be impacted by: – action on zinc transporter (Zip and ZnT) on the release and uptake, respectively, of zinc ions from intracellular stores (1) or the extracellular medium (2) or by modulating the buffering of Zn2+by metallothioneins (3)(41).

Figure 6

Table 3. Glycemic phenotpye of ZnT8 null mouse lines

Figure 7

Fig. 5. (Colour online) Possible sites of action of Zn2+ on insulin signalling. PTB1B, protein tyrosine phosphatase 1B; SHC, Src-homology-2-containing; SOS, Son-of-sevenless; MEK, MAP/ERK kinase; MAP, mitogen-activated protein kinase; IRS, insulin receptor substrate; P1-3K, phosphatidylinositol 3’ kinase; PTEN, phosphatase and tensin homologue; PDK, phosphoinositide-dependent kinase; aPKC, atypical protein kinase C; GSK, glycogen synthase kinase; PPI, protein phosphatase inhibitor-1; GLUT4, glucose transporter family member 4.