Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus...

156
Role of the type II diabetes- associated gene SLC30A8 in the pancreatic α-cell Antonia Solomou Imperial College London Department of Medicine Division of Diabetes, Endocrinology & Metabolism Section of Cell Biology

Transcript of Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus...

Page 1: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

Role of the type II diabetes-

associated gene SLC30A8 in

the pancreatic α-cell

Antonia Solomou

Imperial College London

Department of Medicine

Division of Diabetes, Endocrinology & Metabolism

Section of Cell Biology

Page 2: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

2

DECLARATION OF ORIGINALITY I declare that the following work presented in this thesis is original, except where

indicated by special reference in the text, and that no part of the dissertation has

been submitted for any other degree.

Some of the results contained here, have been presented at conference and

seminars and have been submitted to the scientific literature for publication.

COPYRIGHT DECLARATION

The copyright of this thesis rests with the author and is made available under a

Creative Commons Attribution Non-Commercial No Derivatives licence. Researchers

are free to copy, distribute or transmit the thesis on the condition that they attribute

it, that they do not use it for commercial purposes and that they do not alter,

transform or build upon it. For any reuse or redistribution, researchers must make

clear to others the licence terms of this work.

Page 3: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

3

Acknowledgements

First, I would like to thank Prof Rutter for giving me this opportunity to work in the

lab and earn a PhD. His knowledge and guidance were much appreciated. Secondly, I

would like to thank Dr Hodson for his patience and invaluable help throughout my

PhD, especially in the microscope room.

Life in the lab would have been a lot more boring without Matthew, Ryan, Sophie,

Nick and Natalie. Thanks for all the laughter and the excessive amounts of biscuits

we shared. To all the other lovely people in the group, many of whom have helped

me throughout the years; it has been great working in such a pleasant and relaxed

environment.

Last but certainly not least, I would like to thank Tom, my love and future husband,

who has been there for me right from the start with his love and support whenever I

needed it.

Page 4: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

4

Abstract

The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely

restricted to α- and β-cells, endocrine cells of the pancreatic islet. Genome-wide

association studies have revealed >90 loci associated with Type 2 diabetes. The first

such study identified a specific single nucleotide polymorphism that results in an

amino acid substitution and a transporter with reduced activity. This risk variant is

associated with increased Type 2 diabetes risk. More recently, rare loss-of-function

variants were found to be protective. Here, we aimed to investigate the role of

SLC30A8/ZnT8 in the regulation of glucagon secretion. ZnT8 was selectively deleted

in the α-cell by crossing mice bearing floxed alleles at exon 1 with mice carrying a Cre

recombinase transgene under the control of the preproglucagon promoter.

Additionally, these mice were crossed to Rosa26 RFP mice for identification of α-

cells. Fluorescence-activated sorting of RFP+ cells revealed that recombination at the

RFP locus occurred in ~30% of α-cells. ZnT8 was deleted in ~50% of these labelled

cells achieving a total of ~15% deletion in the whole α-cell population. Glucose

tolerance and insulin sensitivity were normal, though during hypoglycaemic clamps

female KO mice required lower glucose infusion rates and had enhanced glucagon

secretion compared to control mice. Similarly, isolated islets from KO mice released

significantly more glucagon at 1mM glucose. In addition, inducible overexpression of

ZnT8 led to a reciprocal decrease of glucagon secretion from isolated islets at 1mM

glucose. Hypoglycaemic clamps confirmed the finding of lowered glucagon release

following ZnT8 overexpression. Cytosolic Ca2+ responses to low glucose were

comparable in WT and KO cells. Both cytoplasmic and granular free Zn2+ levels were

significantly reduced in the KO α-cells compared to control cells yet there were no

changes in the gene expression levels of other ZnT famlily members. Thus it appears

that ZnT8 is important in the α-cell for appropriate responses to hypoglycaemia and

Zn2+ homeostasis.

Page 5: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

5

Table of contents

DECLARATION OF ORIGINALITY / COPYRIGHT DECLARATION .................................................. 2

Acknowledgements ................................................................................................................... 3

Abstract ..................................................................................................................................... 4

Table of contents ....................................................................................................................... 5

List of figures ........................................................................................................................... 10

Abbreviations .......................................................................................................................... 11

Chapter 1 - Introduction ......................................................................................................... 14

1.1 The pancreas ................................................................................................................. 15

1.2 Islets of Langerhans ....................................................................................................... 15

1.2.1 Islet architecture .................................................................................................... 15

1.2.2 Islet microvasculature ............................................................................................ 16

1.2.3 Islet innervation ...................................................................................................... 17

1.3 Glucagon ........................................................................................................................ 19

1.3.1 Historical perspectives ........................................................................................... 19

1.3.2 Pancreatic α-cell development ............................................................................... 19

1.3.3 Preproglucagon – transcriptional control and processing ..................................... 20

1.3.4 Glucagon receptor .................................................................................................. 24

1.3.5 Physiological actions of glucagon ........................................................................... 24

1.4 Glucagon secretion ........................................................................................................ 25

1.4.2 Regulation by glucose: paracrine or intrinsic? ....................................................... 27

1.4.3 Kinetics of exocytosis.............................................................................................. 28

1.4.4 Stimulation of exocytosis ....................................................................................... 29

1.5 Metabolism of the α-cell ............................................................................................... 30

1.5.1 Glucose ................................................................................................................... 30

1.5.2 Amino acids ............................................................................................................ 32

1.5.3 Fatty acids ............................................................................................................... 33

1.6 Regulators of glucagon secretion – autocrine, paracrine and endocrine ..................... 33

1.6.1 Insulin ..................................................................................................................... 34

1.6.2 Zinc ......................................................................................................................... 35

1.6.3 Somatostatin .......................................................................................................... 35

1.6.4 GABA ....................................................................................................................... 36

Page 6: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

6

1.6.5 Glutamate ............................................................................................................... 36

1.6.6 Glucagon-like-peptide (GLP-1) ............................................................................... 37

1.6.7 ATP .......................................................................................................................... 38

1.6.8 Glucagon ................................................................................................................. 38

1.6.9 Ghrelin .................................................................................................................... 39

1.7 Neural regulation of glucagon secretion ....................................................................... 39

1.8 Diabetes ......................................................................................................................... 40

1.8.1 Type 1 diabetes ...................................................................................................... 41

1.8.2 Type 2 diabetes ...................................................................................................... 41

1.8.3 Monogenic diabetes ............................................................................................... 42

1.9 Genetics of T2D ............................................................................................................. 42

1.9.1 Single nucleotide polymorphisms (SNPs) ............................................................... 43

1.9.2 Genome wide association studies (GWAS) ............................................................ 44

1.9.3 SLC30A8 and T2D risk ............................................................................................. 45

1.10 The α-cell and glucagon secretion in T2D ................................................................... 47

1.11 Zn2+ in diabetes ............................................................................................................ 48

1.12 Cellular Zn2+ homeostasis ............................................................................................ 49

1.12.1 Metallothioneins .................................................................................................. 49

1.12.2 ZIP transporters .................................................................................................... 50

1.12.3 ZnT transporters ................................................................................................... 50

1.13 ZnT8 ............................................................................................................................. 51

1.13.1 ZnT8 and the β-cell; studies using mouse knockout lines .................................... 54

1.13.2 ZnT8 and the α-cell so far ..................................................................................... 56

1.14 Targeting glucagon for T2D treatment ........................................................................ 56

Aims of the thesis ................................................................................................................ 57

Chapter 2 – Materials and Methods ....................................................................................... 58

2.1 In vivo animal work ........................................................................................................ 59

2.1.1 Mice maintenance and generation ........................................................................ 59

2.1.2 Generation of transgenic mice ............................................................................... 59

2.1.3 Glucose tolerance tests .......................................................................................... 59

2.1.5 Insulin tolerance tests ............................................................................................ 59

2.1.6 Fasting plasma glucagon measurements ............................................................... 60

2.1.7 Streptozotocin treatments ..................................................................................... 60

2.1.8 Hypoglycaemic clamps ........................................................................................... 60

Page 7: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

7

2.2 In vitro experiments ...................................................................................................... 60

2.2.1 Islet preparation ..................................................................................................... 60

2.2.2 Islet dissociation ..................................................................................................... 61

2.2.3 Glucagon secretion ................................................................................................. 61

2.2.4 Ca2+ imaging with Fluo-2......................................................................................... 62

2.2.5 Zn2+ imaging with FluoZin-3 .................................................................................... 62

2.2.6 Granular Zn2+ imaging with Zinpyr-4 ...................................................................... 63

2.2.7 Mammalian cell culture .......................................................................................... 63

2.2.8 Zn2+ imaging with eCALWY-4: cells ......................................................................... 64

2.4.4 Amplification and purification of adenoviral eCALWY4 ......................................... 65

2.4.5 Virus titration.......................................................................................................... 65

2.2.9 Immunostaining: cells ............................................................................................. 66

2.2.10 Immunostaining: islets ......................................................................................... 66

2.2.11 Immunostaining: slices ......................................................................................... 67

2.2.12 Quantification of pancreatic glucagon content .................................................... 67

2.2.13 Electron microscopy ............................................................................................. 68

2.3 Molecular biology .......................................................................................................... 68

2.3.1 Genotyping ............................................................................................................. 68

2.3.2 RNA extraction ........................................................................................................ 68

2.3.3 cDNA conversion .................................................................................................... 69

2.3.4 Real Time PCR ......................................................................................................... 69

2.3.5 Fluorescence activated cell sorting (FACS) ............................................................. 69

2.3.6 FACS analysis .......................................................................................................... 70

2.4 Cloning and virus production ........................................................................................ 70

2.4.1 Competent bacteria ................................................................................................ 70

2.4.2 Transfection ............................................................................................................ 71

2.4.3 Bacterial transformation ........................................................................................ 71

2.5 Statistical analysis .......................................................................................................... 71

2.6 Primers .......................................................................................................................... 73

2.7 Antibodies ...................................................................................................................... 75

Chapter 3 - ZnT8 deletion from a subpopulation of α-cells enhances hypoglycaemia-induced

glucagon secretion .................................................................................................................. 76

3.1 Introduction ................................................................................................................... 77

3.2 Aims ............................................................................................................................... 78

Page 8: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

8

3.3 Results ........................................................................................................................... 79

3.3.1 Expression of ZnT8 in the mouse α-cell .................................................................. 79

3.3.2 Generation, breeding and phenotyping of mice null for ZnT8 in the α-cell ........... 80

3.3.3 Recombination efficiency and ZnT8 deletion ......................................................... 82

3.3.4 Body weight and growth ........................................................................................ 84

3.3.5 Glucose homeostasis .............................................................................................. 86

3.3.6 Insulin sensitivity .................................................................................................... 88

3.3.7 Streptozotocin treatment ....................................................................................... 89

3.3.8 Hypoglycaemic clamps ........................................................................................... 89

3.4 Discussion ...................................................................................................................... 91

3.4.1 Recombination in the α-cell ................................................................................... 91

3.4.2 Glucose homeostasis and insulin sensitivity .......................................................... 92

3.4.3 Response to hypoglycaemia in vivo ........................................................................ 93

3.5 Conclusion ..................................................................................................................... 93

Chapter 4 - ZnT8 is necessary in the α-cell for normal cytosolic and granular Zn2+ homeostasis

................................................................................................................................................. 94

4.1 Introduction ................................................................................................................... 95

4.2 Aims ............................................................................................................................... 97

4.3 Results ........................................................................................................................... 98

4.3.1 Impact of α-cell ZnT8 deletion on glucagon secretion in vitro ............................... 98

4.3.2 Calcium responses .................................................................................................. 98

4.3.3 α-cell area ............................................................................................................. 100

4.3.4 Intracellular Zn2+ imaging ..................................................................................... 100

4.3.5 Impact of α-cell ZnT8 deletion on secretory granule Zn2+ content ...................... 105

4.3.6 Gene expression analysis ...................................................................................... 105

4.3.7 Glucagon content ................................................................................................. 107

4.4 Discussion .................................................................................................................... 109

4.4.1 Glucagon secretion ............................................................................................... 109

Zn2+ homeostasis in the α-cell ....................................................................................... 109

4.4.2 Absence of effects of ZnT8 inactivation in α-cells on glucose-regulated Ca2+ fluxes

....................................................................................................................................... 110

4.5 Conclusion ................................................................................................................... 111

Chapter 5 - Overexpression of human ZnT8 in the murine pancreatic α-cell leads to reduced

glucagon release and impaired responses to hypoglycaemia ............................................. 112

5.1 Introduction ................................................................................................................. 113

Page 9: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

9

5.2 Aims ............................................................................................................................. 114

5.3 Results ......................................................................................................................... 115

5.3.1 Generation and genotyping of αZnT8Tg mice ...................................................... 115

5.3.2 Expression of human ZnT8 in the islet ................................................................. 116

5.3.3 Intraperitoneal glucose tolerance was unaltered in α-ZnT8Tg mice ................... 117

5.3.4 Intraperitoneal insulin tolerance tests (ITTs) ....................................................... 118

5.3.5 Hypoglycaemic clamps ......................................................................................... 118

5.3.6 Glucagon secretion is impaired in islets isolated from αZnT8Tg mice ................. 119

5.4 Discussion .................................................................................................................... 122

5.5 Conclusion ................................................................................................................... 123

Chapter 6 – Final Discussion .................................................................................................. 124

6.1 Investigating the role of ZnT8 in the pancreatic α-cell through mouse molecular

genetics ............................................................................................................................. 125

6.2 ZnT8 in the pancreatic α-cell ....................................................................................... 126

6.3 ZnT8 and Zn2+ homeostasis in the α-cell ..................................................................... 131

6.4 ZnT8, SLC30A8 and T2D risk ........................................................................................ 132

6.5 Final conclusions .......................................................................................................... 136

6.6 Future directions ......................................................................................................... 137

Chapter 7 - References .......................................................................................................... 138

Publications ........................................................................................................................... 156

Page 10: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

10

List of figures Figure 1.1 – Mouse pancreas, islets, α- and β-cells ................................................................. 18

Figure 1.2 – A simplified model for the role of transcription factors in endocrine

compartment differentiation .................................................................................................. 21

Figure 1.3 – Structure of proglucagon and processing............................................................ 23

Figure 1.4 Schematic model for KATP channel dependent glucagon secretion. ....................... 31

Figure 1.5 – Subcellular localization and direction of transport of the ZnT and ZIP Zn2+

transporter families. ................................................................................................................ 52

Figure 2.1 – eCALWY sensor design......................................................................................... 65

Figure 3.1 – ZnT8 expression in the α-cell ............................................................................... 79

Figure 3.2 – Generation of αZnT8 KO mice and genotyping by PCR ....................................... 81

Figure 3.3 – Breeding strategy ................................................................................................ 82

Figure 3.4 – Recombination in α-cells ..................................................................................... 83

Figure 3.5– ZnT8 deletion in α-cells. ....................................................................................... 83

Figure 3.6 – FACS analysis of WT and KO islets ....................................................................... 85

Figure 3.7 – Immunostaining of iGluCre islets ........................................................................ 86

Figure 3.8 – Weight measurements ........................................................................................ 87

Figure 3.9 – Glucose homeostasis ........................................................................................... 87

Figure 3.10 – Fasting glucose and glucagon ............................................................................ 88

Figure 3.11 – Insulin sensitivity ............................................................................................... 88

Figure 3.12 – Streptozotocin treatments ................................................................................ 89

Figure 3.13 – Hypoglycaemic hyperinsulinaemic clamps ........................................................ 90

Figure 4.1 - Glucagon secretion in vitro................................................................................... 98

Figure 4.2 – Low glucose-induced calcium responses. ............................................................ 99

Figure 4.3 – Pancreatic histology .......................................................................................... 100

Figure 4.4 – Zn2+ imaging with eCALWY-4 ............................................................................. 101

Figure 4.5 – Islet expression of adenoviral Zn2+ probe .......................................................... 102

Figure 4.6 – AAV-eCALWY-4 .................................................................................................. 103

Figure 4.7 – Intracellular Zn2+ imaging .................................................................................. 104

Figure 4.8 – Granular free Zn2+ imaging ................................................................................ 106

Figure 4.9 – Gene expression analysis ................................................................................... 107

Figure 4.10 – Glucagon content measurements ................................................................... 108

Figure 5.1 – Genotyping by PCR ............................................................................................ 115

Figure 5.2 – Human ZnT8 expression in isolated islets ......................................................... 116

Figure 5.3 – Hum an ZnT8 expression in purified α-cells ...................................................... 117

Figure 5.4 – Glucose tolerance in αZnT8Tg mice .................................................................. 117

Figure 5.5 – Insulin sensitivity ............................................................................................... 119

Figure 5.6 – Elevated glucose infusion rates in αZnT8 KO mice during hypoglycaemic

hyperinsulinaemic clamps ..................................................................................................... 120

Figure 5.7 - Glucagon secretion from αZnT8Tg islets is impaired in vitro ............................. 120

Figure 6.1 – Hypothetical model reconciling data on the effect of Zn2+ on glucagon secretion

and our mouse models .......................................................................................................... 129

Figure 6.2 – Hypothetical model reconciling data for rare and common SLC30A8 variants in

the β-cell ................................................................................................................................ 134

Page 11: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

11

Abbreviations

Arg arginine

Arx Aristaless Related Homeobox

BSA bovine serum albumin

DIO diet induced obesity

DMEM Dulbecco’s modified Eagle’s medium

FBP1 fructose-1,6-bisphosphatase

FRET fluorescence resonance energy transfer

F(1,6)P2 fructose-2,6-bisphosphate

GABA γ-aminobutyric acid

cAMP cyclic AMP

CcgR glucagon receptor

DMSO dimethyl sulfoxide

FACS fluorescence activated cell sorting

FBS fetal bovine serum

FRET förster resonance energy transfer

GIP gastric inhibitory polypeptide

GRPP glicentin-related polypeptide

GLP-1 glucagon-like peptide 1

GRPP glicentin-related polypeptide

GWAS genome wide association studies

G6PC glucose-6-phosphatase

HTRF homogeneous time resolved fluorescence

HVA high voltage activated

Page 12: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

12

IPGTT intraperitoneal glucose tolerance test

ITT insulin tolerance test

KATP ATP-sensitive K+ channel

LB luria broth

LepR leptin receptor

MOI multiplicity of infection

MT metallothionein

Ngn3 Neurogenin 3

Pax-4 Paired box-4

PBS phosphate buffered saline

Pdx-1 pancreatic and duodenal homeobox 1

PCR polymerase chain reaction

PCK2 phosphoenolpyruvate carboxykinase

Pcsk2 Proprotein Convertase Subtilisin/Kexin Type 2

PFA paraformaldehyde

PI3K phosphatidylinositol 3-kinase

PKA Protein Kinase A

PPG preproglucagon

RFP red fluorescent protein

RRP readily releasable pool

tTA tetracycline controlled trancactivator

rtTA reverse tetracycline controlled transactivator

SNP single nucleotide polymorphism

Tet tetracycline

Page 13: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

13

TPEN N,N,N′,N′-tetrakis(2-pyridylmethyl) ethylenediamine

Trp tryptophan

T1D type I diabetes

T2D type II diabetes

VIP vasoactive intestinal neuropeptide

VMH ventromedial hypothalamus

ZIP Zrt-, Irt proteins

ZnT zinc transporter

Page 14: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

14

Chapter 1

Introduction

Page 15: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

15

1.1 The pancreas

The pancreas is an essential organ in nutrient metabolism and serves as both an

exocrine and endocrine gland [1]. The exocrine pancreas, composed chiefly of acinar

cells, is involved in nutrient digestion through the production and secretion of

digestive enzymes into the intestine through the pancreatic duct. These include

enzymes such as carboxypeptidase-A, trypsin and amylase [2]. The endocrine

pancreas functions largely independently from the exocrine compartment and is

responsible for the maintenance of glucose homeostasis by secreting hormones into

the bloodstream. Endocrine cells cluster together to form the islets of Langerhans,

micro-organs that can be found scattered throughout the pancreas between the

exocrine tissue (Figure 1.1A) [3].

During embryogenesis, the pancreas is formed by fusion of a dorsal and ventral bud

created by evagination of the foregut. In the mouse the dorsal bud develops on

embryonic day 9.5 at the same time as the first differentiated glucagon expressing α-

cells develop [4]. Insulin-positive cells appear the next day, often simultaneously

producing glucagon [5]. By day 14, much larger numbers of β-cells can be seen and

the endocrine cells begin to arrange into small clusters. By the following day δ-cells

expressing somatostatin are present. Then, briefly before birth, the pancreatic

polypeptide (PP)-producing cells develop and the endocrine cells initiate the

formation of the highly organized islets of Langerhans [6].

1.2 Islets of Langerhans

1.2.1 Islet architecture

The islets of Langerhans account for 1-2% of total pancreatic mass only (~2000 in

rodents and 1 million per pancreas in humans). They tend to be spherical, 50-200 μm

in diameter and range in size from 1000 to 10000 cells [7]. There are several

different endocrine cell types in islets which work co-operatively to regulate glucose

metabolism: α-cells produce glucagon, β-cells insulin, δ-cells somatostatin, PP cells

pancreatic polypeptide. Additionally, a separate population of islet cells has been

identified the ε-cells which produce ghrelin, a hormone secreted mainly by the

Page 16: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

16

stomach to influence feeding behaviour [8]. The two main hormones are insulin and

glucagon which are released reciprocally in response to glucose fluctuations to keep

its levels within a relatively small range. Islets thus constitute a multihormonal

micro-organ, its sole purpose the maintenance of glucose homeostasis.

Rodent islets have a specific islet architecture where the vastly abundant β-cells are

clustered within the islet core surrounded by a thin layer of predominantly α-cells

along with the other endocrine cells (Figure 1.1B). This anatomical subdivision is not

present in human islets as the β-cells intermingle with α- and δ- cells throughout the

islet [9]. This lack of segregation of the different cell types in human islets has been

reported by various groups [9, 10] and could be suggestive of stronger paracrine

interactions compared to rodent islets. The cellular composition of islets also differs

between rodents and humans. In mouse islets [9] the β- cells make up the vast

majority of the islet reaching 70-80%, α-cells 15-20%, δ-cells 5-10% and the PP cells

<1%. In human islets the proportion of β-cells is only 50-60%, α-cells 30-40% and δ-

cells ~10% [9]. α- and β-cells are virtually indistinguishable from each other except

when achieving the high resolution of an electron microscope where the subtle

differences in the appearance of their secretory granules can help identify them

(Figure 1.1C,D). Imaged by conventional transmission electron microscopy β-cell

granules have a halo surrounding their dense core while α-cell granules appear

dense throughout [11].

1.2.2 Islet microvasculature

Pancreatic islets are highly vascularised with a dense network of sinusoidal

capillaries [12]. These capillaries have about 10 times more fenestrations than those

in the exocrine pancreas [13] that allow direct exchanges between the blood and

islet cells. In the rat, islets receive 10-15% of total blood flowing to the pancreas [14].

This direct interface allows the islet to respond rapidly to changing glucose

concentrations and other nutrients with hormonal release into the circulation. The

direction of blood flow within the islet is still an issue of debate. It has been

proposed that, in the rodent islet, blood flows initially through β-cell areas before

reaching non-β cell populations or that alternatively non-β-cells receive blood first

Page 17: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

17

before β-cells [15]. In human islets no anatomical subdivision was observed so that

all endocrine cell types were randomly distributed along blood vessels [9]. Such

findings suggest the theory that there is no specific direction of islet perfusion so

that each cell type could influence different cell types not excluding its own cell type

[15].

1.2.3 Islet innervation

Islets are extensively innervated to allow autonomic regulation over hormone

release. Both sympathetic (adrenergic) and parasympathetic (cholinergic) fibres

reach deeply into the islet. The sympathetic system is stimulated by hypoglycaemia,

exercise or stress, releasing noradrenaline [16] into the islet along with the

neuropeptides galanin [17] and neuropeptide Y [18]. Activation of the sympathetic

system triggers the secretion of glucagon with adrenaline and noradrenaline acting

via the α1- and β-adrenoceptors [19]. In contrast, adrenaline and the sympathetic

neuropeptides are inhibitors of both insulin [20] and somatostatin secretion [21].

The parasympathetic system which is activated during hyperglycaemic conditions

releases the neurotransmitter acetylcholine [22] and neuropeptides such as

vasoactive intestinal polypeptide [23]. Parasympathetic activation exerts an overall

stimulatory effect on the islet inducing secretion of insulin, glucagon, somatostatin

and PP [24, 25]. The majority of research on islet innervation has been focused on

the sympathetic system while the contribution of the parasympathetic mediators to

enhancing glucagon secretion is still unclear. Sensory neurons releasing γ-

aminobutyric acid (GABA) [26] and other neurotransmitters such as nitric oxide [27],

cholecystokinin [28], ATP [29], glutamate [30], dopamine and serotonin [31] have

also been identified in islets.

In humans, however, cholinergic innervation of islets is sparse [32]. In contrast to

mouse islets, acetylcholine acts as a paracrine signal released from α-cells affecting

the surrounding endocrine cells of the islet rather than being primarily a neural

signal.

Page 18: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

18

Figure 1.1 – Mouse pancreas, islets, α- and β-cells. (A) Representative image of a mouse

pancreas acquired by widefield microscopy. The green and red dots represent islets while

the rest is the exocrine pancreas. (B) Representative image of an islet stained for insulin

(green) and glucagon (red). Scale bar 50 µm. Transmission electron microscopy image of a

mouse α- cell (C) and β-cell (D). Note the difference in the structure of the dense core

vesicles between the two cell types. Scale bar 1 µm.

A

C B

D

Page 19: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

19

1.3 Glucagon

1.3.1 Historical perspectives

The history of glucagon started at the same time as that of insulin. When Banting

and Best tested their primary pancreatic extracts on pancreatectomised dogs, in

1921, they noticed that hypoglycaemia caused by insulin was followed by a mild

hyperglycaemia which they believed was a consequence of adrenaline release [6].

Murlin et al [33] discovered that injecting these pancreatic extracts caused an

increase in blood glucose in both healthy and diabetic dogs. They attributed this

effect to a glucogenic substance, and this was called glucagon [33]. Then in 1948

Sutherland and de Duve demonstrated, through a series of experiments, that the α-

cells of the pancreas were supplying the glucagon [34]. A few years later Foa et al

performed cross-circulation experiments in anaesthetised dogs which showed that

glucagon release is stimulated by hypoglycaemia [35]. Advances in glucagon

purification [36], and identification of its amino acid composition [37] subsequently

made studies into its physiology possible. In 1961, Unger et al developed glucagon

antibodies and a glucagon radio-immuno assay for assessing glucagon levels [38] and

in 1962 glucagon was localised to α- cells by immunofluorescence staining [39].

1.3.2 Pancreatic α-cell development

Multiple transcription factors have been discovered with temporally and spatially

limited expression during pancreas formation (Figure 1.2). During pancreatic

development, pancreatic duodenum homeobox-1 (Pdx-1) is expressed to promote

the development of the gland [40]. The transcription factor Neurogenin 3 (Ngn3), is

necessary to stimulate the next step, i.e. for progenitor cells to follow an endocrine

path. Ngn3 deletion in mice leads to an absence of pancreatic endocrine cells [41].

Neuro D is then needed for proliferation and is expressed in all endocrine cells [42].

Aristaless Related Homeobox (Arx) is responsible for correct endocrine development

as Arx-deleted mice exhibit an early onset loss of mature α-cells and at the same

time experience elevated numbers of β- and δ-cells [43]. Therefore, Arx is essential

for α-cell fate and to restrict β- and δ-cells. This function in endocrine commitment is

Page 20: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

20

opposite to that of Paired box-4 (Pax-4) since there is a build-up of Pax-4 and Arx

transcripts when Arx and Pax-4 are deleted in mice, respectively [43]. This

antagonism between the two transcription factors for islet cell differentiation

determines the earliest stage of progression into α- or β-cells and is dependent on

the levels of each transcript [6].

Brain specific homeobox-4 (Brn-4) activates the glucagon gene [44] alongside MafB

which is solely expressed in the α-cell and further promotes cell type specific

transcription of glucagon [45]. Pax-6 is also necessary for differentiation into α-cells

determined by studies on mice which, when lacking Pax6, developed either very few

or no glucagon-expressing cells [46].

1.3.3 Preproglucagon – transcriptional control and processing

The proglucagon gene is expressed in pancreatic α-cells, intestinal L- cells, and in the

brain, primarily the hypothalamus and thalamus, hindbrain and olfactory bulb [6].

The preproglucagon gene was discovered and isolated from humans and rodents in

the 1980s. Following molecular cloning the isolation of hamster preproglucagon

cDNA [47] and the human glucagon gene [48] revealed the existence of three

homologous sequences, glucagon, GLP-1 and GLP-2 [48, 49].

These are joined together by two intervening peptides IP-1 and IP-2 and are

downstream of the signal peptide glicentin-related polypeptide (GRPP) at the N

terminus.

Mammalian preproglucagon is composed of 179-180 amino acids. The three

peptides are bound to each other by a pair of basic amino acids, potential sites for

proteolytic cleavage by proteases involved in hormone maturation [50].

Page 21: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

21

Figure 1.2 – A simplified model for the role of transcription factors in endocrine

compartment differentiation. The position of each transcription factor indicates its time of

expression or its main functional role. The most important transcription factors in the

differentiation of α- and β-cells are shown. Some are expressed at several steps but they are

only depicted once for simplicity. From Gromada et al [6].

Page 22: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

22

The hormone that will eventually be produced from preproglucagon depends on the

presence of specific proteases known as prohormone convertases. Proprotein

Convertase Subtilisin/Kexin Type 2 (Pcsk2) is exclusively found in the α-cell and is

responsible for processing proglucagon into active 29 amino acid-long glucagon,

GRPP and IP-1, peptides with unknown functions, and the major proglucagon

fragment (MPGF) of inactive GLP-1, IP-2 and GLP-2. Conversely, in the L cell the

presence of Pcsk1/3 and the lack of Pcsk2 lead to the production of GLP-1 and GLP-2

alongside IP-2 and glicentin, the N-terminal fragment that includes unprocessed

glucagon which is processed to GRPP and oxyntomodulin (Figure 1.3).

The promoter of the preproglucagon gene has six key elements upstream the

transcription initiation site. These sequences of 20-40 bp are called G1, G2, G3, G4,

CRE (cAMP response element) and ISE (intestinal specific element). The G1 sequence

allows α-cell specific expression of the gene [51] while G2-G4 are islet specific

enhancers [52]. The CRE sequence is responsible for transcription in response to

cAMP [53] and ISE confers expression specifically in the L-cell [54]. Pax-6 is the main

transcription factor during pancreas development through transactivation of G1 [55],

while Pax-4 inhibits glucagon expression by preventing Pax-6 transactivation [56].

Pdx-1 acts as a repressor of glucagon transcription [57] and is important for islet cell

differentiation being an important transactivator of the preproinsulin gene [58].

Insulin has been shown to act as an inhibitor of glucagon expression in α-cells by

affecting transcription rates through an insulin-responsive DNA element in the

proximal 5’ flanking region of the G3 promoter sequence [59]. Later studies proved

that in addition to the proximal DNA regions distal promoter elements are involved

in the insulin mediated inhibition of glucagon expression [60].

Page 23: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

23

Figure 1.3 – Structure of proglucagon and processing. Top: structure of proglucagon. At the

end of each peptide there are pairs of dibasic amino acids (KK, RR, KK where K is lysine, R

arginine) and these are potential cleavage sites by prohormone convertases. Bottom left:

Proglucagon processing occurring in the α-cell. The peptides are GRPP, glucagon, IP1 and

MPGF which can be partially processed to GLP-1. Bottom right: Processing of proglucagon in

the L cell. The products are glicentin, truncated GLP-1, IP2 and GLP-2. Glicentin can be

partially processed to GRPP and oxyntomodulin. Adapted with permission from Furuta M et

al., J. Biol. Chem., 2001 [61].

GRPP

glucagon

IP1

GRPP glucagon GLP-1 GLP-2

KR KR KR RR RR KK

COOH

IP1 IP2

KR

GLP-1

MPGF

PC2 PC1/3

GRPP

GLP-1

IP2

Oxyntomodulin

GLP-2

Proglucagon

α-cell L cell

Glicentin

Page 24: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

24

1.3.4 Glucagon receptor

The rodent glucagon receptor (GCGR), composed of 485 amino acids is a member of

the secretin-glucagon receptor class II family and is G protein-coupled receptor [62].

The binding of glucagon is coupled to a Gαs type GTP-binding G protein that in turn

causes the activation of adenylate cyclase and subsequently production of cAMP and

Protein Kinase A (PKA) activation. Glucagon binding can also induce the

phospholipase C/inositol pathway through Gq proteins, mobilising intracellular Ca2+

stores [62]. The glucagon receptor is found in several tissues being most abundant in

the liver and kidney and to a lesser extent in the endocrine pancreas, heart, cerebral

cortex, adipose tissue and smooth muscle regulating various processes in these

tissues to modulate glucose homeostasis [63].

Genetic polymorphisms in the GcgR have been found to be linked to Type 2 diabetes

(T2D). For example a missense mutation in exon 2 of the GcgR gene that causes an

amino acid change is associated with T2D in French populations. The mutated

receptor appears to have a decreased affinity for glucagon and lowered cAMP

responses after stimulation [64]. However, the mutation was not associated with

T2D in other populations [65]. Glucagon receptor null mice have lower glucose levels

although they are not hypoglycaemic but display α-cell hypertrophy and severe

hyperglucagonaemia [66].

1.3.5 Physiological actions of glucagon

The body has developed several ways to defend against hypoglycaemia and the likely

harmful effects that could ensue, particularly in the brain, which relies heavily on a

steady supply of glucose as a fuel. Such defences include a reduction in insulin

secretion and elevated release of adrenaline and glucagon [67]. Glucagon has a key

role in the response to hypoglycaemia, its primary action being in the liver where the

ratio of insulin/glucagon regulates various aspects of hepatic metabolism. Thus,

glucagon induces gluconeogenesis and glycogenolysis resulting in an increase of

hepatic glucose production for release into the bloodstream and subsequent supply

of glucose to the rest of the body and very importantly to the brain [68]. It

Page 25: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

25

simultaneously decreases glycolysis and glycogenesis which would use up glucose.

Through the highly selective glucagon receptor, glucagon mediates its actions in the

liver mainly by activating the adenylyl cyclase and the PKA pathway. Through this

mechanism, glucagon elevates gluconeogenesis by the upregulation of enzymes

including glucose-6-phosphatase (G6PC) and phosphoenolpyruvate carboxykinase

(PCK2) via induction of the cAMP response element-binding protein (CREB) and

Peroxisome proliferator-activated receptor gamma 1 (PPARγ1) [69, 70]. Another

important player is coactivator transducer of regulated CREB activity 2 (TORC2).

Glucagon causes the dephosphorylation and subsequent translocation of TORC2 to

the nucleus where it enhances CREB-dependent transcription [71]. PCK2, G6PC and

fructose-1,6-bisphosphatase (FBP1) are important in the rate of gluconeogenesis.

PCK2 is responsible for the conversion of oxaloacetate into phosphoenolpyruvate

and G6PC catalyses glucose transition of glucose-6-phosphate to glucose. FBP1

mediates production of fructose-6-phosphate from fructose-1,6-bisphosphate

(F(1,6)P2). The activity of this enzyme is controlled by glucagon as it reduces the

levels of fructose-2,6-bisphosphate (F(2,6)P2), its allosteric inhibitor [72]. This

reduction of F(2,6)P2 in turn reduces phosphofructokinase-1 activity, decreasing the

flux through glycolysis. Glycolysis is also inhibited by glucagon at the pyruvate kinase

step [73].

The direction of glycogen metabolism is determined primarily by the action of

glycogen synthase (GS) and glycogen phosphorylase (GP). Glucagon mediates GP

phosphorylation and activation and at the same time inactivates GS by

phosphorylation thus causing glycogenolysis [74, 75]. Additionally, in aid of elevating

glucose levels glucagon stimulates the uptake of amino acids by the liver for

gluconeogenesis [76].

1.4 Glucagon secretion

Each α-cell contains ~ 7500 secretory granules [77]. These granules have a similar

sized dense core to β-cells with a core diameter of ~230 nm. They are round and

have a dark electrondense core but lack the characteristic halo of insulin containing

Page 26: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

26

granules making them smaller in overall diameter [78]. About 1% of these granules

are docked beneath the plasma membrane and make up the readily releasable pool

(RRP) [77].

1.4.1 Ion channels in the α-cell

α-cells differ sharply from β-cells in their response to glucose, generating action

potentials in the absence of sugar or at sub-physiological levels. Activation of

voltage-gated Na+ and Ca2+ currents is needed for the initiation of the action

potential and K+ currents are necessary for spike repolarization. Several types of ion

channels are found in the α-cell. There are four types of K+ selective channels; the

ATP sensitive K+ channel (KATP), the delayed rectifying K+ channel, a G-protein gated

K+ channel stimulated by somatostatin and a transient K+ channel. Also identified are

four types of voltage gated Ca2+ channels (T, N, R, L), a Na+ channel and GABA

receptor chloride channels [6].

KATP channels have a key function in coupling cell metabolism to electrical activity in

a variety of tissues [79]. In the β-cell, during resting conditions, the basal activity of

the channel sustains the negative membrane potential until glucose metabolism

leads to a concentration-dependent inhibition of the channel, due to a change in the

ATP/ADP ratio within the cell. KATP closure causes membrane depolarization and the

subsequent opening of voltage-dependent Ca2+ channels. Ca2+ influx then ensues

followed by Ca2+ dependent exocytosis [6]. The KATP channel in α-cells is highly

sensitive to ATP and thus more sensitive to inhibition than in the β-cell requiring

concentrations as low as 0.05 mM to cause maximal stimulation of activity[6].

Voltage-dependent K+ channels are necessary for repolarization of the α-cell action

potential, while Na+ channels are required for the production of an action potential.

Delayed rectifying K+ currents have been detected in mouse, guinea pig and in rat α-

cells [77, 80]. This current is sensitive to tetraethylammonium which can decrease

the amplitude of the outward current by >90%. Its application to rat α-cells led to

enhanced glucagon secretion due to elevated membrane depolarisation [81].

Page 27: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

27

The Na+ channels are tetrodotoxin-sensitive and are functional at potentials ranging

from positive to -30 mV in mouse α-cells. Their function is essential for generation of

action potentials. Tetrodotoxin suppresses glucagon secretion from rodents to a

level similar to that seen by high glucose [82].

The voltage-gated Ca2+ channels the α-cell is equipped with belong to three

subfamilies which are: 1) the L-type high voltage-activated (HVA) channels, 2) P/Q-,

N-, and R- type HVA Ca2+ channels 3) the low voltage activated T-type channels [80].

The opening of the T-type channels takes place at membrane potential as low as -65

mV while for HVA channel activity depolarisations of -40 to -30 mV and above are

needed. The different requirements for voltage values indicate that the low and high

voltage channels could serve separate functions in action potential generation. T-

type channels are active around the action potential threshold and it has been

suggested that they are involved in action potential initiation. HVA channels open at

membrane potentials near the rising phase of the action potential and are of greater

capacity accounting for the majority of the Ca2+ entry necessary for glucagon release

[6].

1.4.2 Regulation by glucose: paracrine or intrinsic?

The question of whether the inhibition of glucagon secretion by glucose occurs due

to the release of factors from neighbouring cells or whether α-cells have the intrinsic

ability to sense changes in glucose concentrations is a matter of debate. Insulin and

glucagon secretion being reciprocally regulated by glucose prompted the speculation

that glucagon secretion is controlled in a paracrine manner by insulin which is known

as the ‘switch-off hypothesis’ [83]. Such a hypothesis would easily explain the

hypersecretion of glucagon observed in diabetes, it being a result of low insulin

release, and why this is abolished with the aid of exogenous insulin. Variations to the

concept propose factors released with insulin from secretory granules could be

contributing to or be responsible for the suppressive effect of high glucose on α-cells

[84].

Page 28: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

28

The issue with this hypothesis is that the α-cells are maximally inhibited at glucose

levels below the threshold for β-cell activation and insulin secretion or somatostatin

[85]. Studies indicate that isolated single α-cells still respond to glucose by lowering

intracellular Ca2+ concentrations and suppression of glucagon release [86, 87]. The

suppressive effect of glucose was preserved when insulin signalling was blocked

using wortmannin [88]. Additionally there is evidence of vagal control of glucagon

secretion [89], yet following denervation the α-cell response to hypoglycaemia is

intact [90]. Such findings suggest that although the α-cells might be under paracrine

regulation they also possess an intrinsic ability to sense glucose.

This mechanism remains poorly defined but research on mice lacking functional KATP

channels, which have an impaired regulation of glucagon secretion by glucose,

suggests that KATP channels are involved in some way [91]. Mice lacking the KATP

channel subunit Sur1, in hypoglycaemic conditions, release similar amounts of

glucagon to WT islets exposed to maximally inhibitory glucose concentrations and

glucagon secretion cannot be inhibited further when the glucose concentration is

increased. Additionally, in WT mouse islets, the suppression of glucagon release by

high glucose can be reversed by low concentrations of diazoxide, a KATP channel

activator [82]. Human α-cells similarly express KATP channels and administering the

KATP channel blocker tolbutamide results in depolarisation suggesting that these

channels are active in the α-cell and have a repolarizing role [92].

1.4.3 Kinetics of exocytosis

Exocytosis of glucagon from the α-cells is Ca2+ dependent just like secretion of insulin

from β-cells, but is slightly faster. The difference is a result of the quantity of

granules in the RRP, approximately 120 in the α-cell compared to ~60 in the β-cell,

and the rate at which the RRP is refilled, ~ 20 granules/sec in the α-cell and 5-10

granules/sec in the β-cell [93]. Additionally, there is a robust secondary acceleration

of the exocytotic rate in the α-cell following the emptying of the RRP which could be

due to fast Ca2+ dependent translocation of secretory granules deeper into the cell to

Page 29: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

29

the release sites. The maximum exocytosis rate of 750 granules/sec during

depolarisation is completed within 20 msec and is similar to that observed in β-cells

which is ~700 granules/sec and finishes in 15 msec [93].

Considering such findings it appears that the two cell types share a comparable

organization of Ca2+-dependent exocytosis, but suggest a higher number of release

sites is present in α-cells. The large Ca2+ currents of the smaller α-cells, resulting in

larger and less localised Ca2+ transients, could at least partly account for the

differences between the two cell types. Such features allow the α-cell to act rapidly

and secrete glucagon during short periods where glucose is in demand such in cases

of intense physical activity [93].

1.4.4 Stimulation of exocytosis

α-cells are electrically excitable and use electrical signals to couple any changes in

glucose levels to changes in glucagon output. They fire action potentials in the

absence of glucose, possessing a specific set of channels to generate currents of Na+

and Ca2+, and this electrical activity in turn produces Ca2+ signals and glucagon

release.

A key role for KATP channels is the widely most accepted hypothesis for glucose

sensing and glucagon secretion (Figure 1.4). In the mouse, at low glucose

concentrations, the function of the KATP channels is to keep the membrane potential

at ~-60 mV which is the threshold for action potential initiation. At such a voltage, T-

type Ca2+ channels become active bringing the membrane potential within a range

that allows activation of Na+ and HVA L-type and N-type Ca2+ channels which open at

membrane potentials more positive than -30mV. This leads to the regenerative

opening of these channels and action potential generation [85]. The entry of Ca2+

through N-type channels is what stimulates glucagon release. Repolarisation of

action potentials is achieved by the efflux of K+ through voltage dependent K+

channels.

Page 30: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

30

An increase in glucose leads to a rise in cytosolic ATP/ADP levels which causes the

closure of the KATP channel. Subsequently the α-cell is depolarised but, unlike in the

β-cell the depolarisation suppresses glucagon release. The α-cell action potential

depends on T-type Ca2+ channels and Na+ channels and at the membrane potential

following sustained depolarisation the channels become inactivated thus inhibiting

electrical activity, Ca2+ fluxes and glucagon release [85, 94]. Therefore glucagon

secretion is primarily supported by an intermediate KATP channel activity which keeps

the membrane potential within a range that can promote regenerative action

potentials [85]. A similar mechanism has been suggested for glucagon secretion by

human islets [85].

The above mechanism has been questioned since it has been reported that glucose

may have a hyperpolarising instead of a depolarising effect [95, 96]. An alternative

model has been proposed of glucose suppressing glucagon release independently of

KATP channels by inhibition of a depolarizing Ca2+ store-operated current [95, 97].

1.5 Metabolism of the α-cell

1.5.1 Glucose

It has been well established that glucose triggers monophasic suppression of

glucagon secretion from islets. Based on dose-response experiments the α-cell

appears to be more sensitive to glucose compared to the β-cell [98]. The threshold

for glucagon inhibition is 2-3 mM glucose which is lower than the β-cell threshold of

4-5 mM which acts as a stimulus for insulin release. Additionally half-maximal

inhibition of glucagon is at 3-6 mM, while half-maximal stimulation of insulin

happens at 8-10 mM. Maximal suppression of glucagon secretion is achieved at 6-10

mM glucose, but for maximal stimulation of insulin release the glucose

concentration has to be above 20 mM [98, 99]. This difference in sensitivity makes it

unlikely that increased insulin secretion as glucose concentrations rise is responsible

for the inhibition of glucagon release through a paracrine mechanism. Of note, other

Page 31: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

31

Figure 1.4 Schematic model for KATP channel dependent glucagon secretion. Under low

glucose conditions (top) the ATP/ADP ratio is low so KATP channels still maintain a low level of

activity. This causes moderate depolarisation which activates Na+ channels resulting in

action potential generation and a Ca2+ flux through Ca2+ channels. This leads to Ca2+-

dependent exocytosis of glucagon. At high glucose levels (bottom) there is increased entry of

glucose through Glut2 transporters which increases the ATP/ADP ratio and causes the

closure of KATP channels. This in turn inactivates Na+ channels so that there is no action

potential generation and no exocytosis. Adapted from Zhang et al [100].

glut 1

glut 1

Page 32: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

32

related sugars and glucose metabolites have the same inhibitory impact on glucagon

secretion and the suppressive effect is based on the ability of the α-cell to

metabolize the sugar [101]. This observations is in agreement with the fact that

suppression of glucagon secretion by glucose is prevented by inhibitors of cellular

metabolism [102].

α-cells like β-cells express the high-Km glucokinase whose activity in the α-cell is

similar to that in the β-cell and is thought to have a glucose sensing function [103].

Additionally there is expression of the low-Km hexokinase II which is responsible for

~50% of glucose phosphorylation in α-cells compared to less than 10% in β-cells

[103]. This flux may be have a role in sensing very low glucose levels as it reaches

saturation at 1mM glucose [103].

The glucose transporter Glut1 is expressed in the α-cell instead of the more efficient

(but low affinity) Glut2 found in the rodent and human β-cells [104]. The lack of

Glut2 and the low levels of Glut1 mean that the uptake rate of glucose by the α-cells

is 10 times less than that by β-cells yet is it still 10 fold greater than the metabolic

rate so that transport is not a rate limiting factor in terms of glucose metabolism

[104]. The rate of glucose utilization i.e. glycolysis is similar in α-cells and β-cells

[104]. However glucose oxidation, i.e. oxidative phosphorylation, in α-cells reaches

only 10% of that in the β-cell [105]. Non quantitative, time-resolved measurements

of cytosolic ATP levels using a luminometric method have shown that glucose

increases induce a rapid increase in ATP [88]. Measuring ATP and ADP

concentrations biochemically in pure α-cells and β-cells indicate that the ATP/ADP

ratio in α-cells at 1mM glucose is the same as that in β-cells subjected to 6 mM

glucose [105].

1.5.2 Amino acids

Amino acids have a stimulatory effect on glucagon secretion, with an action on the

α-cell which is more potent than that on the β-cell [106]. The capacity of each amino

acid to stimulate glucagon secretion differs, with arginine being the most effective,

alanine and glutamine being potent but less so [86]. Some amino acids have little or

Page 33: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

33

even no effect individually, but can potentiate the stimulation of glucagon secretion

by other amino acids [86]. Amino acids that overall have a small impact on insulin

secretion have a key role in the regulation of glucagon secretion under physiological

conditions. Since glucose enhances insulin responses to amino acids while

dampening glucagon release, amino acids in combination with glucose play a crucial

part in the differential stimulation of the cell types. Insulin release after a protein

meal would lead to hypoglycaemia if it was not accompanied by a stimulation of

glucagon secretion [6].

1.5.3 Fatty acids

Fatty acids influence glucagon release but the extent varies among different species

and experimental conditions [107]. The initial observation was that they have a

suppressive effect on glucagon release [108], but later studies showed that short

term exposure actually induces its secretion [109]. The short term stimulatory effect

of fatty acids on mouse islets is depended on chain length, the longer the chain the

higher the glucagon release. Spatial configuration is also important, with trans fatty

acids being more potent than cis isomers. The level of saturation similarly impacted

efficacy with saturated fatty acids being more effective than unsaturated [110].

Palmitate can increase exocytosis by enhancing Ca2+ influx through L-type Ca2+

channels and by reducing somatostatin release from δ-cells by 50% thus alleviating

the α-cell from its inhibition [111]. The long term effect of palmitate and oleate on α-

TC1.6 cells was to enhance glucagon secretion in a dose dependent manner but

suppress proliferation [112]. Consistent with these findings, long term exposure of

rat islets led to an enhancement in glucagon secretion, a reduction in glucagon

content likely a result of increased release since glucagon gene expression was

unaltered [113].

1.6 Regulators of glucagon secretion – autocrine, paracrine and endocrine

While the regulation of insulin secretion is well understood, the regulation of

glucagon secretion is a lot less clear with many hypotheses on how various factors

might affect it in a paracrine manner, being released from neighbouring cells, or how

Page 34: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

34

it might be centrally controlled through innervation. These are discussed in more

detail below.

1.6.1 Insulin

β-cell secretory granules contain a high concentration of crystalline hexameric insulin

(~118 mM in human and 74 mM in rat primary cells [114]) and once released the

insoluble hexamers separate into the active monomeric form. It can thus be found at

a high local concentration in the islet with the potential to influence adjacent cells.

Insulin has long been regarded as the most likely contender of all the β-cell products

to have an inhibitory effect on secretion from α-cells [115]. In diabetic and non-

diabetic humans intravenous infusion of insulin, while maintaining normoglycaemia,

resulted in a reduction in glucagon secretion [116]. The effect of insulin on glucagon

secretion appears to be conserved across species. Thus exogenous administration of

insulin has been shown to inhibit glucagon release from perfused streptozotocin-

treated rat pancreas [117], streptozotocin-treated hamsters [118] and guinea pig

islets [119], alloxan-treated diabetic dogs [120], rat pancreas and isolated α-cells

including isolated rodent islets, but only at low glucose [121], and clonal mouse α-

cells [122]. Conversely, perfusion of isolated rat [123] and human [124] pancreata

with anti-insulin antibodies at 5.5 and 11.5 mM glucose respectively stimulated

glucagon release.

Insulin receptor transcripts are highly expressed in rat α-cells [121] and in clonal

mouse and hamster α-cells [122]. Insulin reduced glucagon secretion from both

mouse and hamster cell lines [122]. Deletion of the insulin receptor in mice

specifically in the α-cell caused mild glucose intolerance, hyperglycaemia and

hyperglucagonaemia in the fed state as well as an elevated response to arginine

[125]. Various groups have provided some insight into the mechanism by which

insulin acts in the α-cell. Evidence from rat α-cells suggests insulin (0.1 µM) has a

hyperpolarising effect on the α-cell by increasing KATP channel activity thus inhibiting

electrical activity and glucagon secretion [121]. In another study on primary mouse

α-cells, 3 µM insulin significantly reduced the sensitivity of KATP channels to inhibition

by ATP shown to be mediated via the phosphatidylinositol 3-kinase (PI3K) signalling

Page 35: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

35

pathway [126]. Further to its effect on KATP channels, insulin, via the activation of a

PI3K-Akt dependent pathway, can cause the translocation of GABA-A receptors to the

plasma membrane leading to an increased response to GABA secreted by β-cells,

membrane hyperpolarization and suppression of glucagon release [127].

1.6.2 Zinc

In its crystalline form in secretory granules, insulin is centrally coordinated to two

Zn2+ ions to form hexamers [128]. Zn2+ levels in these granules reach ~20 mM [129]

and upon exocytosis the ions dissociate and are released into the microvasculature

or interstitial space [121] achieving a high local concentration with the potential to

act as mediators of paracrine signalling. An inhibitory effect of Zn2+ on glucagon

secretion was first observed when the Zn2+ chelator Ca2+-EDTA allowed stimulation

of secretion by the normally inhibitory mitochondrial substrate monomethyl-

succinate in the perfused rat pancreas while insulin secretion was unchanged [84]. In

support of this finding, Zn2+ was found to inhibit pyruvate- or glucose-induced

glucagon secretion from isolated rat α-cells by reversibly activating KATP channels and

thus reducing electrical activity and glucagon release [121]. Another study

demonstrated this inhibitory effect of Zn2+ treatment on glucagon secretion in both a

glucagon producing mouse cell line and intact mouse islets [130]. However, this was

not replicated in a similar study which showed no effect of Zn2+ on glucagon

secretion from mouse islets and a murine α-cell line [88].

1.6.3 Somatostatin

Somatostatin is widely distributed throughout the body being produced by both the

central and peripheral nervous system but also peripheral organs such as gut,

kidneys, adrenal and thyroid glands and the δ-cells of the pancreas [131].

Immunocytochemical studies revealed that out of the five somatostatin receptor

(SSTR) subtypes SSTR2 is highly expressed in α-cells while SSTR1 and SSTR5 are

expressed in β-cells in human islets [132]. In rodents SSTR2 and SSTR5 predominate

in the α-cells and β-cells respectively [133, 134]. Somatostatin has been established

Page 36: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

36

as an inhibitor of both insulin and glucagon secretion [134, 135], released in the islet

is response to glucose with a threshold concentration of 4-5 mM.

Three mechanisms have so far been described as to how activation of the

somatostatin receptors causes the inhibition of glucagon release. Electrophysiology

data indicate that activation of G protein-coupled K+ channels by somatostatin leads

to hyperpolarisation of the cell membrane in isolated mouse and rat α-cells so that

electrical activity is inhibited [136, 137]. Somatostatin has also been found to

stimulate a Gαi2 protein-dependent pathway followed by localised activation of the

phosphatase calcineurin that in turn causes depriming of docked secretory granules

in primary rat cells. This effect did not involve a change in Ca2+ levels and

hyperpolarisation [138]. Additionally, somatostatin binding to its receptor inhibits

adenylyl cyclase activity and thus decreases cAMP levels and PKA induction of

glucagon release [139].

1.6.4 GABA

GABA is a major inhibitory neurotransmitter in the central nervous system but it is

also produced by β-cells which package it into microvesicles similar to synaptic

vesicles for release in a controlled manner [140]. GABA was shown to cause

hyperpolarization of rat α-cells upon activation of the GABAA receptor. The use of an

antagonist to explore the effect of endogenous GABA was found to increase

glucagon release from rat islets twofold at 1mM glucose and eliminated the

inhibition imposed by 20 mM glucose. However, the levels of secretion were still

lower than at low glucose conditions [141]. Exogenous application of GABA exerted

an inhibitory effect on arginine-induced glucagon release in guinea pig [142] and

mouse islets [143], primary rat α-cells [81] and the mouse line αTC6 [144].

1.6.5 Glutamate

Islet endocrine cells appear to have glutamatergic signalling elements with the ability

to make, release and respond to glutamate, the major excitatory neurotransmitter in

the CNS [145]. Glutamate is produced in β-cell mitochondria and in the cytosol, likely

Page 37: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

37

accumulating in insulin granules with the possibility of acting as an intercellular

signalling agent following exocytosis [146]. Several of the subunits of the large

glutamate receptor family have been found to be expressed in the different types of

islet cells. In the α-cell both ionotropic receptors, AMPA and kainite subtypes, have

been identified [147] as well as metabotropic receptors [148]. The ionotropic

receptors act as Na+ channels whose activation regulates plasma membrane

potential and thus glucagon release. In neurons, the G-protein coupled metabotropic

receptors, modify cAMP levels and are involved in autocrine feedback inhibition of

glutamate signalling. Binding of glutamate to AMPA or kainite receptors in a

hypoglycaemic setting stimulated glucagon secretion from the perfused rat pancreas

[149]. Another study showed a modest inhibition of glucagon secretion from isolated

islets and the perfused rat pancreas following administration of GABA, an effect

believed to involve ionotropic receptors [143].

Vesicular glutamate transporter subtypes 1 and 2 have been identified in α-cell

glucagon granules and a concomitant elevation in glucagon and glutamate release

has been observed from rat islets in low glucose conditions [150]. Consequently,

glutamate co-released with glucagon might have a positive autocrine effect on

glucagon secretion. In contrast, the activation of α-cell metabotropic receptors leads

to suppression of glucagon release from rat islets under hypoglycaemic conditions

[151]. The evidently opposing effects of the two different types of glutamate

receptors following activation on glucagon release augment the level of complexity

regarding intra-islet glutamate signalling.

1.6.6 Glucagon-like-peptide (GLP-1)

GLP-1, an incretin secreted by L-cells in the small intestine following a meal [152],

primarily acts to enhance glucose-induced insulin release through GLP-1 receptors

on β-cells and via activation of the autonomic nervous system. GLP-1 also acts on the

α-cell to inhibit glucagon release although the expression of GLP-1 receptors in α-

cells is <0.2% of that in β-cells [153]. The suppressive ability of GLP-1 was observed in

vivo and in the perfused pancreas. One group concluded that its inhibitory effect

Page 38: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

38

must be paracrine since in single isolated cells GLP-1 actually stimulated glucagon

release through interaction with G-protein couple receptors activating adenylate

cyclase which leads to cAMP production [154]. However, similar research on both

intact and isolated cells concluded that suppression of glucagon release by GLP-1 is

PKA-dependent, glucose-independent and does not involve paracrine mechanisms.

The incretin appears to inhibit not by affecting electrical activity rather by selective

inhibition of N-type Ca2+ channels and thus exocytosis [153].

1.6.7 ATP

ATP is another potential paracrine inhibitor of glucagon release. It can be found in

synaptic vesicles of islet nerve endings but also in insulin secretory granules [155].

Using immunocytochemistry α-cells were shown to express the purinergic receptors

including the P2 receptor subtype P2Y1 and the P1 receptors P1A1 and P1A2A [156].

Applying ATP extracellularly diminished or reduced the frequency of calcium

oscillations observed in α-cells at low glucose both in intact mouse islets and isolated

cells. This effect appears to be directly through ATP and not due to its hydrolysed

products and most likely involve P2Y1 and A1 receptors [156]. Previous findings from

perfused rat pancreata [157] and rat islets [158] contradict the inhibitory effect of

ATP on mouse islets since adenosine and ATP respectively stimulated glucagon

release. Such contrasting results could be due to species differences.

1.6.8 Glucagon

Glucagon has a stimulatory effect on insulin and somatostatin release from the

perfused pancreas of both rats and humans [124, 159]. It also has an autocrine role

having a positive feedback effect by binding to glucagon receptors expressed in α-

cells, elevating cAMP levels and exocytosis from both rat and mouse α-cells [160].

Experiments on mice with an inability to produce glucagon due to the absence of the

processing enzyme PCSK2 [161] or missing the glucagon receptor [162] suggest that

the lack of this autocrine control by glucagon or the loss of glucagon activity in target

tissues causes α-cell hyperplasia. Reducing glucagon receptor activity specifically in

the mouse liver led to enhanced glucagon secretion in vivo, but there was no

Page 39: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

39

increase in α-cell mass indicating that the trigger stimulating hyperplasia was most

likely an autocrine effect and not due to the lowering of glucagon signalling in the

liver [163].

1.6.9 Ghrelin

Ghrelin is a growth hormone-releasing peptide synthesised by the stomach in fasting

conditions. It is also found in the human and rodent islets, released from a separate

cell population distinct from all other cell populations (α, β, δ, PP) of the islet [8]. It

appears that ghrelin-positive cells are more abundant in the foetal and neonatal

stage, diminishing in numbers at the adult stage, and seen mainly in the periphery of

the islet [164]. The ghrelin receptor is widely expressed in the islet and was observed

to colocalize with α-cells [165]. Ghrelin had a stimulatory effect on glucagon

secretion from isolated mouse islets, but no such effect was observed in vivo [166].

Additionally, and more recently ghrelin, was found to increase glucagon release from

mouse αTC1 and InRG9 guinea pig cell lines but also from isolated islets in a dose

dependent manner [167].

1.7 Neural regulation of glucagon secretion

As described earlier (Section 1.2.3) islets are amply innervated by sympathetic and

parasympathetic nerves to ensure rapid prevention or correction of hypoglycaemia

for protection of the brain [168]. Reductions in arterial glucose levels are sensed in

various areas of the brain, carotid bodies and the hepatic portal vein. A specific

region in the brain, the ventromedial hypothalamus (VMH) has been identified as an

important player in sensing falling blood glucose concentrations and also in

commencing counterregulatory measures against hypoglycaemia [169, 170]. There

are three levels of autonomic control over the α-cell, sympathetic and

parasympathetic nerves and circulating adrenaline, all of which are activated by

hypoglycaemia.

Stimulation of the sympathetic and parasympathetic arms of the nervous system is

achieved by glucose-sensing neurons in the VMH. The glucose sensing process is

Page 40: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

40

thought to involve KATP channels since mice lacking a component of these channels

have a permanently raised firing rate and an extremely defective glucagon counter-

regulatory response [89]. Activating the channels in the VMH neurons

pharmacologically leads to enhanced adrenaline and glucagon responses during

hypoglycaemia [171].

Adrenaline is a strong stimulator of glucagon release in rodent islets caused by an

increase in cAMP levels, an effect which can be reproduced by using the adenylate

cyclase activator forskolin [153]. However, adrenaline appears to have no such

ability on glucagon release in human islets and similarly forskolin had no effect either

[172]. A stronger counterregulatory response by α-cells in response to adrenaline

could be a likely explanation as to why mice manage insulin-induced hypoglycaemia

better than humans [172].

Parasympathetic and sympathetic nerve terminals may contain stores of

neurotransmitter including acetylcholine and noradrenaline as well as several

neuropeptides. Once released these may either stimulate or suppress glucagon

depending on the specific peptide found in each terminal [63]. Cholinergic activation

through muscarinic receptors leads to cytoplasmic Ca2+ mobilization, increasing

glucagon release, as observed in isolated mouse α-cells and also in vivo [173]. The

same effect was detected after noradrenaline administration, which caused

adenylate cyclase activation and subsequently increased intracellular Ca2+

concentration and glucagon secretion [174]. Additionally, neuropeptides including

vasoactive intestinal neuropeptide (VIP) and gastric inhibitory polypeptide (GIP) have

the ability to induce glucagon secretion and can be stored in parasympathetic nerve

terminals. Galanin and neuropeptide Y is found in sympathetic terminals [168].

1.8 Diabetes

In the UK alone the number of people diagnosed with diabetes has reached 3.5

million and this figure is expected to rise to 5 million by 2025 making diabetes one of

the biggest health problems. The majority of cases (~90%) are patients with Type 2

Page 41: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

41

diabetes (T2D). Globally, the prevalence of diabetes is expected to increase from 171

million in 2000 to 366 million by 2030 [175]. Diabetes is a consequence of the loss of

glycaemic control and islet dysfunction is a characteristic of both Type 1 (T1D) and 2,

as defined in Sections 1.8.1 and 1.8.2 below. It was 40 years ago that the bihormonal

hypothesis to explain the diabetes pathophysiology was first proposed [176]. This

hypothesis postulated that the disease was caused by an insulin deficit or resistance

accompanied by absolute or relative glucagon excess which results in an elevated

rate of glucose production by the liver surpassing glucose utilisation and thus leading

to hyperglycaemia. Insulin deficiency is the major feature of diabetes and results in

chronic hyperglycaemia with glucose levels above 7 mM which can lead to long-term

health issues such as kidney failure, neuropathy, blindness and cardiovascular

disease [177].

1.8.1 Type 1 diabetes

Type I is caused by the T cell mediated auto-immune destruction of β-cells. In T1D

patients, islets become infiltrated by CD8+ T cells which appear to be exclusively

specific towards islet autoantigens [178]. This usually occurs in people with a genetic

susceptibility and is believed to be initiated by environmental stressors such as a

viral infection [179]. The major susceptibility locus is located at the HLA class II genes

[180]. Patients normally have islet autoantibodies and the infiltration of islets by

immune cells results in an overall reduction of 70-90 % of β-cell mass at clinical onset

with most islets having lost their β-cell element [181].

1.8.2 Type 2 diabetes

This form of disease has reached epidemic proportions in various parts of the world,

a result of adverse environmental components, particularly the increased

accessibility to high caloric food accompanied by a sedentary lifestyle and the lack of

sufficient levels of physical activity [182]. The heritability of T2D is well established,

and variations of specific genetic loci interact with such environmental factors which

predispose to disease and have an impact on the development of the diabetic

phenotype [183]. Loss of glycaemic control in this type of diabetes is a combined

Page 42: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

42

result of defective insulin secretion from β-cells and insulin resistance of insulin

target tissues such as muscle, liver and adipose tissue [184]. Development of obesity

is an important risk factor foretelling the eventual development of insulin resistance

[185] which, in an individual with a genetic predisposition to β-cell impairment,

causes a change in glucose tolerance. In the initial stages of T2D patients maintain

their glucose tolerance by an increase in functional β-cell mass and hypersecretion of

insulin as a compensatory mechanism for insulin resistance. However, as the disease

advances β-cells are no longer able to secrete enough insulin to reach the levels

required for peripheral tissues and that is when glucose intolerance begins [184].

Hyperglycaemia in the fasting and fed state, a consequence of dysregulated glucose

production by the liver and diminished glucose clearance, is attributed to ineffective

insulin action. A role for glucagon in the pathogenesis of T2D and a bihormonal axis

of the disease has long been suspected and multiple studies now support this view

[186].

1.8.3 Monogenic diabetes

The prevalence of monogenic diabetes is thought to be at 2-5% of the diabetic

patient population [187]. It represents a heterogeneous set of disorders caused by

faults in single genes. The monogenic variants are categorized based on the age of

onset, neonatal diabetes appearing before six months of age and maturity onset

diabetes of the young (MODY) showing signs before 25 years of age, with an

autosomal dominant type of inheritance and insulin independence [188]. MODY can

be caused by mutations, deletions or other abnormalities in any of at least twelve

different genes resulting in intrinsic β-cell dysfunction. These genes, which include

glucokinase [189], the transcription factors HNF-4α [190], HNF-1α [191], HNF-1β

[192], IPF1 [193] and NeuroD1 [194], are all expressed in β-cells.

1.9 Genetics of T2D

Although the global epidemic of T2D is largely driven by lifestyle and nutritional

changes, the blending of environmental factors with susceptible genetic traits

promotes the development of the disease [195]. For T2D patients diagnosed in

Page 43: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

43

adulthood, the concomitant action of such susceptibility alleles and several

combinations of frequent variants at many loci could lead to detrimental effects

upon interaction with the aforementioned predisposing environmental components

[182].

The concept of a genetic influence on T2D is now commonly acknowledged. A study

on young patients with T2D revealed that the majority of the parents were either

diabetic or at least glucose intolerant. Similarly, there was high prevalence of

diabetes or glucose intolerance in the siblings, reaching about 70% [196]. The high

prevalence of diabetes in particular ethnic groups provides further evidence of a

genetic component [197]. Studies on monozygotic twins revealed that when one was

diagnosed with diabetes 48 out of the 53 pairs were concordant, with the majority of

the second twin in each pair becoming diabetic within five years of the first [198].

The ability to scan the whole genome for variation and do so using high throughput

technology made it possible to search for genetic variation to explain phenotypic

diversity and disease risk in genome-wide association studies (GWAS) (Section 1.9.2).

Assaying single nucleotide polymorphisms (SNP) to study the association of genetic

variation and disease has now become the norm and can give clues into the genetic

architecture contributing to complex disease [199].

1.9.1 Single nucleotide polymorphisms (SNPs)

SNPs are single base pair changes in the DNA sequence that are abundant in the

human genome [200]. In the case of genetic studies, SNPs are normally used as

markers of a part of the genome with the vast majority of them having no or minor

effect on the biology of an individual. However, SNPs can have a functional impact, a

consequence of causing amino acid changes, or, in the case of SNPS in non-coding

(intergenic or intronic regions), altering transcript stability or transcription factor

binding ability [201]. SNPs are the most frequent type of genetic variation in humans

[202] occurring on average every 300 base pairs (bp). Normally, there are two forms

of each SNP (one per haploid chromosome) so that in any given population, two

naturally occurring base pairs will likely exist for each specific SNP location with each

Page 44: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

44

individual being either heterozygous or homozygous for an SNP. The major allele is

the most common variant of the SNP, while the minor the one with the lower

frequency.

1.9.2 Genome wide association studies (GWAS)

Although linkage analysis and candidate gene approaches have in the past been

successful in revealing familial genetic defects with strong effects and have allowed

the identification of genes that cause MODY, these approaches are not effective in

identifying common genetic variants which have a relatively moderate effect on

disease risk [199].

In GWAS studies, the genome of patients and non-patients is compared to find SNPs

that are associated with disease. Through the use of microarrays a staggering

number of 500 000 to a million SNPs are genotyped during each study, allowing for a

much faster discovery of risk variants linked to T2D. These kinds of studies can lead

to the discovery of several gene variants, each individual one having a small impact

on disease risk. When a particular SNP is found associated with disease, the locus is

typically annotated with the name of the nearest gene. Despite this, it should not be

automatically assumed that the variant involved is the molecular defect that

accounts for the disease phenotype or that the closest gene is necessarily implicated

with disease. The annotation merely marks a specific part of the genome that

contains the variant which could actually be functioning at a distance for example

regulating gene expression at a distant site [199].

The first GWAS for T2D was performed on a French case-control cohort, comprising

661 T2D cases and 614 healthy controls testing a total of 392,935 SNPs for

association with the disease. This study led to the identification of novel and

reproducible association signals at the SLC30A8 and HHEX loci [203] and at the same

time confirmed the already known association at TCF7L2 [203]. Shortly afterwards, a

study on an Icelandic population, confirmed the loci just mentioned and identified a

Page 45: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

45

new association that of a CDKAL1 variant with an odds ratio (OR) of 1.20 [204].

Further replication was provided by Zeggini et al [205].

An OR is a statistical term in this case used to quantify how strongly a specific allele

is associated with disease. An OR significantly greater than 1 would mean an allele is

associated with disease. The strongest association was mapped to a variant in the

transcription factor TCF7L2 gene which had been previously identified in Icelandic

and Danish populations [206]. Each copy of the risk allele carries a high odds ratio for

T2D of 1.4-1.5 [207]. Inheritance of the risk allele can also be used to predict the

possibility of transitioning from prediabetic conditions to T2D [208].

Currently the model of T2D susceptibility in adulthood is that, being a complex

metabolic disease, it is probably the consequence of environmental issues on

predisposed individuals bearing a mixture of increased risk genetic variants. It is

believed that each risk allele, on its own, has a moderate effect with odds ratios of

1.10-15 [182].

Ever since the initial investigation [203] several others have undertaken such studies,

confirming previously identified disease signals or discovering new T2D associated

variants [205, 209]. Although a lot of the common variant signals revealed through

GWAS are associated with impaired islet function, the majority of the signal from the

studies is in non-coding regions and some in introns [210]. Determining functional

links to specific genes is thus more challenging requiring the deployment of both

mapping approaches [211] and the use of “functional genomics” to test directly the

impact of forced changes in the expression of candidate genes in model systems

(cells, animals etc) [212].

1.9.3 SLC30A8 and T2D risk

One of the first variants to be discovered by GWAS, in the initial study by Sladek et al

[203], was rs13266634, a non-synonymous SNP in SLC30A8 located in a 33 kb linkage

disequilibrium block on chromosome 8, which includes only the 3’ end of the gene

[203]. The SNP causes an amino acid change from tryptophan to arginine at position

Page 46: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

46

325 at the C-terminus of the ZnT8 protein. Exon 13, encoding the last 52 amino

acids of both ZnT8 isoforms contains a total of 6 SNPs, with a second SNP occurring

at position 325, rs16889462, in this case changing the amino acid from an arginine to

glutamine. The gene product ZnT8 is expressed primarily in pancreatic islets [213]

where it is likely to play an important role in islet hormone secretion.

The rs13266634 allele was associated with increased T2D risk. Possession of only one

allele of the risk variant results in an OR of 1.18, while being homozygous, increases

the OR to 1.53 [203]. Subsequent studies performed on T2D patients from different

countries found similar association between SLC30A8 and disease [204, 214]. The

association of this SNP with T2D typically involves comparatively young onset and

leanness, although it was not identified as a cause of MODY following extensive

studies on 53 extended families of patients with early onset T2D [215]. Of note, ZnT8

was identified as a target by autoantibodies in 60-80% of new onset T1D patient

[216] though the risk variants discovered by GWAS do not affect T1D risk.

Studies performed on a German population of T2D patients [217, 218] revealed that

the risk SLC30A8 allele, which is also the major allele in most populations, is

associated with decreased insulin release following intravenous administration of

glucose and defective conversion of proinsulin to insulin during an oral glucose test,

whilst there were no signs of insulin resistance. Another study on a French

population [219] demonstrated that those with the risk allele had higher plasma

glucose levels in fasting conditions and decreased insulin secretion. However, other

studies showed no connection between the SNP and markers of β-cell function [214,

220].

Unexpectedly, a more recent study [221] showed that very rare loss of function

variants of SLC30A8 are actually protective and were associated with decreased T2D

risk. 12 protein truncating variants were identified and carriers of such variants had a

65% reduction in T2D [221]. Such observations conflict with the findings that

expression of a less active (R325) form of ZnT8 [222] results in an elevated risk of

Page 47: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

47

T2D suggesting that a more intricate mechanism is at play regarding the rs13266634

T2D association [223].

1.10 The α-cell and glucagon secretion in T2D

Similar to the β-cell, the α-cell becomes dysfunctional in diabetes. Plasma glucagon is

significantly elevated in cases of uncontrolled T1D and contributes greatly to

ketoacidosis [224]. Glucagon levels are abnormal in T2D as well, although the

changes are more subtle, tending towards enhanced secretion and/or insufficient

inhibition of α-cells in most patients. Such inappropriately high glucagon levels

appear to have a role in hyperglycaemia and impaired glucose regulation [186]. In

conditions of inadequately-controlled diabetes or ketoacidosis, plasma glucagon

levels are very high [224]. Hyperglucagonaemia in the fasted state has been

observed in some but not the entirety of T2D patients with at least moderate

glycaemic control [225].

In the diabetic islet, the α-cell does not respond appropriately to raised plasma

glucose. The suppression of glucagon secretion as seen in healthy individuals by

elevation of glucose levels does not occur in T2D. The response to hyperglycaemia is

either blunted or lacking and plasma glucagon stays abnormally high at similar blood

glucose concentrations [186]. Additionally, stimulation of glucagon release by

normal stimuli including intravenous arginine or protein-rich meals is greater in T2D

compared to healthy individuals [226]. The inability of hypoglycaemia to induce

glucagon secretion is well established in T1D, but whether this is the case in T2D is

more controversial. Several groups have found the counter-regulatory response to

hypoglycaemia by the α-cell to be weakened at least in individuals with long duration

of T2D [227, 228].

Fasting hyperglycaemia in T2D is directly attributed to abnormal hepatic glucose

production and hepatic resistance to suppression of glucose production by insulin is

evident in patients [186]. However, surplus glucagon in the fasting state could also

be a driver for inappropriate levels of glucose production by the liver. Basal glucagon

release is a significant factor in the control of fasting glucose levels in both diabetic

Page 48: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

48

and healthy individuals [225, 229]. Fasting glucagon can be up to 50% higher in some

T2D patients compared to non-diabetics [225].

In order to determine the importance of glucagon in diet-induced and genetic T2D in

a rodent model, as caused by leptin receptor (LepR) ablation, one group studied high

fat diet-induced obese (DIO) mice and LepR KO (LepR-/-) mice with or without

glucagon receptors (GcgR) [230]. GcgR+/+ and GcGR-/- mice were placed on a high

fat diet and, as previously observed, the body weight of the KO mice did not increase

substantially and they did not exhibit hyperinsulinaemia or hyperglycaemia in

contrast to the WT mice. DIO and LepR-/- mice with functional GcgRs were both

hyperinsulinaemic and obese developing mild and severe T2D respectively. LepR-/-

GcgR-/- mice showed no signs of hyperinsulinaemia and hyperglucagonaemia.

Delivering GcgR to the liver through the use of an adenovirus led to the severe

hyperinsulinaemia and hyperglycaemia normally present in LepR-/- mice, while

spontaneous disappearance of the transgene reversed the T2D phenotype. Such

findings indicate that glucagon is essential for hyperglycaemia in T2D and that it

enables diet-induced hyperinsulinaemia. These effects can be prevented by blocking

glucagon action.

1.11 Zn2+ in diabetes

A total of 2-3 g of Zn2+ can be found throughout the human body and this ion is most

abundant in the pancreas which contains 140 μg/g [231] with the highest

concentration within islets. In the plasma total Zn2+ reaches 10-20 μM and its free

concentration is in the nM range [232]. A role for Zn2+ in diabetes pathophysiology

was discovered almost 80 years ago, when a 50 % reduction in pancreatic Zn2+ was

observed in the diabetic compared to non-diabetic cadavers [233]. Then, about 30

years later, studies in rodents revealed that a Zn2+ deficiency through diet led to a

decrease in glucose-induced insulin secretion and insulin levels in the islet [234].

Following that, it was shown that Zn2+ deficiency also causes a reduction in β-cell

granulation [235]. It has been reported that T2D patients exhibit a substantial

reduction in total blood Zn2+ as well as at the cellular level in comparison to healthy

Page 49: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

49

controls accompanied by excessive zincuria [236]. Dietary supplementation

improved glucose handling in human subjects [237].

Research on various genetic mouse models of T2D demonstrated decreased

pancreatic Zn2+ levels [238]. Supplementation of db/db mice with Zn2+ led to

normalization of pancreatic Zn2+ concentration and a lessening of hyperglycaemia

and hyperinsulinaemia [239]. The same experiment in ob/ob mice increased

pancreatic insulin levels and similarly reduced the observed fasting hyperglycaemia

and hyperinsulinaemia as well as the excessively high insulin release in response to

glucose in vitro [238]. The opposite experiment of a Zn2+-deficient diet in db/db mice

caused an exaggeration of fasting hyperglycaemia [239].

1.12 Cellular Zn2+ homeostasis

Zn2+ is very important biologically having catalytic, structural and regulatory roles

[240]. Additionally, it is involved in a plethora of cellular processes and is believed to

be a cofactor for approximately 3000 proteins in the human genome including

enzymes, transcription factors and hormones [241]. Mechanisms that regulate Zn2+

uptake, distribution and efflux are crucial for maintaining homeostasis and thus

cellular function. Intracellular Zn2+ homeostasis is maintained in part through the

activity of Zn2+ binding proteins and a variety of Zn2+ transporters belonging to two

protein families which have opposing function but work in a coordinated way (Figure

1.5).

1.12.1 Metallothioneins

The metallothioneins (MTs) are intracellular Zn2+-binding proteins essential in

regulating Zn2+ uptake distribution, storage and release. They are redox active,

having sulphur donors, controlling Zn2+ homeostasis but also contributing to the

redox state of the cells by releasing Zn2+ in the presence of oxidative stress [242].

The capacity of MTs for intra- and intermolecular Zn2+ exchange means that MTs

keep sufficient Zn2+ accessible in the cell at micromolar concentrations, higher than

the concentration of free Zn2+ in the picomolar range, which would be too low to act

as a reservoir of Zn2+ for Zn2+ proteins [243]. Several SNPs in the MT genes have been

Page 50: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

50

discovered to be associated to T2D following studies done on an Asian population

[244].

1.12.2 ZIP transporters

The ZIP (SLC39A) family of Zn2+ transporters has 14 members (Zip1-14) in mammals.

The majority are predicted to have eight transmembrane domains and a long loop

with a histidine-rich region which is likely to be the Zn2+ -binding domain. They have

a similar predicted topology with both their N and C termini located on the

extracellular surface of the plasma membrane [245]. Their function is the facilitation

of Zn2+ influx from either the extracellular space or within intracellular

compartments into the cytoplasm, thus increasing the cytoplasmic free Zn2+

concentration.

1.12.3 ZnT transporters

The ZnT (SLC30A8) family comprises 10 members, ZnT1-10, and most are predicted

to be made up of six transmembrane domains with cytoplasmic carboxy and amino

termini. Similarly to the Zips, they contain a long histidine-rich loop potentially being

the site of Zn2+ binding in the cytosol [246]. ZnT5 differs in that it has 15

transmembrane domains [247], whilst ZnT6 has a serine rich loop instead [248]. The

majority of ZnTs form homodimers, except ZnT5 and ZnT6 which form heterodimer

responsible for transporting Zn2+ early in the secretory pathway [249]. ZnTs have

been found in cytoplasmic compartments normally associated with the Golgi

complex, endosomes or endoplasmic reticulum [250]. In contrast to Zips they are

efflux transporters, carrying Zn2+ from the cytosol into intracellular compartments or

extracellularly.

Zn2+ binds to the histidine rich region of the protein and the transmembrane

domains form the pore for Zn2+ transport in both transporter families [246]. The

different isoforms vary in their affinity for Zn2+ and also have the ability to transport

other transition metals, for example manganese, cadmium, iron and copper which

Page 51: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

51

may inhibit the transport of Zn2+ [251]. Transcription and translation of at least some

of these transporters appears to be regulated directly by Zn2+ [251].

Some members of the ZIP and ZnT families have ubiquitous expression but others

are tissue specific. For example ZnT1, ZnT5 and Zips 6-10 are found throughout the

body, but ZnT10 is only expressed in the liver and brain [251].

1.13 ZnT8

ZnT8 was first identified as a pancreas-specific ZnT member and cloned by Chimienti

et al in 2004 [213]. It was predicted to adopt the same topology at the other ZnT

proteins, having 6 transmembrane helical domains with a histidine rich region

between helices 4 and 5. These authors showed that only the pancreas had

significant expression of ZnT8 and that although not complete there was a high

degree of colocalization between insulin and EGFP-tagged ZnT8. The protein was

detected in intracellular vesicles and at the plasma membrane. Additionally, using

Zinquin a Zn2+ indicator that localizes to intracellular vesicles [252], ZnT8 was found

to be involved in Zn2+ transfer across the vesicle membrane following transient

expression in HeLa cells which do not normally express the transcript. Since co-

staining for ZnT8 with insulin granules was not complete, ZnT8 could possibly be

localised at other subcellular sites and so it was found to partially colocalize with the

endoplasmic reticulum marker calnexin.

The human SLC30A8 gene is found on chromosome 8, contains eight exons across

37kb and is translated into a 369 amino acid protein. Mouse and rat ZnT8 have been

shown, respectively, to share 80 and 76% identity with the human form of ZnT8 and

are most closely related to ZnT2, ZnT3 and ZnT4 of the ZnT family. The greatest

homology is with ZnT2, reaching 53.5% [253]. Initially ZnT8 expression in the mouse

was found to be largely restricted to the pancreas, specifically in the α- and β-cells of

the islet [130]. Similarly analysis of a large number of human tissues confirmed the

pancreas specific expression of ZnT8. Later reports showed its expression, albeit at

low levels, in human adipose tissue [254], lymphocytes [255] and the cubical

epithelium lining thyroid follicles and the adrenal cortex [256].

Page 52: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

52

Figure 1.5 – Subcellular localization and direction of transport of the ZnT and ZIP Zn2+

transporter families. The direction of Zn2+ transport is shown by arrow for the ZnT (green)

and ZIP proteins. From Myers et al [251].

Page 53: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

53

Upon SDS-PAGE gel analysis ZnT8 migrates as bands of ~ 40 and 90 kDa, indicating

that it may form dimers. X-ray crystallography on the E.Coli Zn2+ transporter, a

bacterial ZnT8 homologue Yiip, showed that it is a homodimer bound in a parallel

orientation by co-ordinating to four Zn2+ [257]. The amino acid at position 325 in the

ZnT8 sequence is predicted to reside at the dimer interface. Nicolson et al [222]

demonstrated that the risk variant has a reduced ability to transport Zn2+ across the

granule membrane, with the position of the changed amino acid believed to be in a

different position to the one mentioned above, in a region near the predicted site of

Zn2+ binding or docking sites to Zn2+-binding proteins. Introducing positive charge

into this area through the presence of arginine instead of tryptophan might possibly

affect the kinetics of Zn2+ transport and the recruitment of Zn2+ binding proteins

[258]. Similar findings, supporting this view, were later made by Kim et al [259] using

radiotracers to monitor Zn2+ fluxes into isolated secretory granules. On the other

hand Weijers [260] has presented an alternative structure prediction suggesting that

the R-W325 replacement has little effect on structure or interactions with Zn2+ or

binding proteins. A resolution of this issue will require crystallisation of the ZnT8

protein itself (or the C-terminal domain) as well as further functional assessment of

the impact of the common polymorphism on activity.

There is not a lot of information on the transcription factors involved in the

regulation of ZnT8 expression. A strong transcriptional enhancer has been

discovered in the second intron of the mouse gene [261] which is highly conserved in

humans [262]. It was found that this binds several transcription factors one of them

being Pdx1. However this enhancer is only active in β-cells and not α-cells. In terms

of factors that could modulate its expression, Zn2+ deprivation, glucose and cytokines

have been shown to reduce Slc30A8 expression in both mouse and rat cell lines [263,

264] though no such effect was observed on human islets after administration of

cytokines [265].

Page 54: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

54

1.13.1 ZnT8 and the β-cell; studies using mouse knockout lines

Zn2+ is essential in the β-cell for the correct processing and storage of insulin in

secretory granules [266]. After translation proinsulin is transferred to the Golgi body

where Zn2+ is abundant and there proinsulin hexamerizes with Zn2+, two ions in each

hexamer. In the immature secretory vesicle, proinsulin in processed into insulin and

insulin crystallization happens in the insulin granule, creating the characteristic

dense core [267].

Thus given this importance of Zn2+ in insulin granules and since ZnT8 was discovered

to be a T2D gene, there was a lot of interest into its role in the β-cell. Functioning

across the granule membrane to transport Zn2+ into the lumen for insulin

crystallization, ZnT8 has the potential of being extremely important for the β-cell.

The role of ZnT8 in the β-cell has been examined with either global or β-cell specific

deleted mouse strains. The results from these studies were controversial, with not

everyone agreeing with the findings of the other, although there were some

concurring points. None of the groups observed any changes in total insulin content,

islet size or in the proportion of cell types [222, 268-271]. In the majority of the

cases, whether global or β-cell specific ablation, the insulin granules appeared to

have been changed morphologically [222, 270-272]. Only one study reported no

changes [269]. Reductions in cellular Zn2+ content were also reported though there

were some differences between groups. One group observed a reduction in Zn2+

content with the use of Zinquin and dithizone yet not with FluoZin-3 [222], perhaps

reflecting the localisation of these dyes to the granules (zinquin and dithizone) and

cytosol (FluoZin-3). On the other hand, Pound et al did demonstrate a significant

decrease in cytosolic free Zn2+ using FluoZin-3 [268, 269]. Additionally another group

only showed a modest reduction in cytosolic Zn2+ when looking at β-cell specific KO

mice versus controls [272] compared to the changes observed in global KO mice

[222]. Of note the fluorescent intracellular probes used above provide uncertain sub-

cellular localisation and may be accumulated into organelles including secretory

granules to different degrees.

Page 55: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

55

Glucose-stimulated insulin secretion from isolated islets was one of the most divisive

findings since two groups described an impairment of secretion [268, 272], two

elevated secretion [222, 271] and two no effect at all [269, 270]. It has been

suggested that the increase in insulin release might be due to the alleviation of

inhibition by Zn2+ which was previously reported [271].

Zn2+ is important for the maturation of the enzymes involved in proinsulin processing

and so proinsulin levels and processing were investigated. Again the results were in

disagreement, with two studies detecting no changes in proinsulin processing [222,

270], two showing a small increase in plasma levels [271, 272] while Pound et al

[269] observed a reduction in plasma proinsulin levels which was, however, affected

by genetic background. Thus it appears that despite the marked changes in granule

morphology the deletion of ZnT8 has minor effects on proinsulin processing per se.

All of the groups above examined the effect of ZnT8 deletion on in vivo glycaemic

parameters. No effect of ZnT8 deletion on insulin sensitivity was reported and no

change in fasting blood glucose except by one group who showed an increase but

only in young mice [222]. Three studies observed no changes in fasting insulin [269]

[271, 273] and only one detected a decrease [222]. However, most studies showed

impaired glucose tolerance during intraperitoneal glucose tolerance tests in young

mice [222, 269-271] with only two reporting such an effect in old mice [222, 268-

271]. Lastly, one group reported impaired glucose tolerance assessed by oral glucose

tolerance test in young mice [269], but not in older ones [272].

Studies in the INS-1 rat β-cell line using an shRNA-mediated approach to eliminate

ZnT8 expression by >90% resulted in reduced uptake of exogenous Zn2+, lowered

insulin content and decreased insulin secretion supporting the view suggesting that

ZnT8 may have a role in insulin release and the control of glycaemia [263].

Page 56: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

56

1.13.2 ZnT8 and the α-cell so far

The presence of ZnT8 in the α-cell has been confirmed by qPCR and by

immunocytochemistry [222]. Just as in the β-cell, colocalization of ZnT8 with

glucagon was only partial with ZnT8 found in other cytosolic locations other than the

secretory granules. The function of ZnT8 in the α-cell has not been extensively

studied. Only one group has previously provided a limited amount of data on a

αZnT8 KO mouse [222]. Intraperitoneal glucose tolerance tests showed that the

absence of ZnT8 had no impact on glucose homeostasis. Similarly, there were no

differences in plasma glucagon levels when compared to control mice [272].

However, these studies were confined to mice of a specific age, and no other

parameters were tested.

There have been some studies done on the effect of overexpression or knockdown

of ZnT8 on αTC1.9 cells. Overexpression of either the R325 or W325 variants in these

cells caused a reduction in glucagon content and a decrease in glucagon secretion by

50%. Knockdown of ZnT8 reciprocally led to an elevation of glucagon mRNA and

increased glucagon secretion by 70% [274].

1.14 Targeting glucagon for T2D treatment

Taking into consideration the scientific evidence supporting glucagon as a

contributing factor in abnormal glucose homeostasis in T2D, targeting glucagon

action could be a potential therapeutic strategy.

Reducing plasma glucagon levels using somatostatin, while insulin levels are kept

constant, was found to normalise the enhanced basal rate of hepatic glucose

production and fasting plasma glucose concentration in hyperglycaemic patients

with T2D [225]. Glucagon receptor antagonists have been developed which, in

leptin-deficient diabetic mice, have been demonstrated to be effective in lowering

the enhanced basal rate of hepatic glucose production and fasting plasma glucose

levels [275].

Page 57: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

57

Orally administered glucagon receptor antagonists have also been produced but the

concern regarding this type of drugs is the ensuing α-cell hyperplasia and the

constant elevation of blood glucagon concentration [276]. There is also evidence,

however, that glucagon receptors are required in the liver to prevent hepatic lipid

accumulation [277], a significant counter-indication in T2D. However, drugs which

cause a reduction in plasma glucose concentration without severe α-cell

hypertrophy and only a modest elevation in blood glucagon levels have been

developed [278]. Two of these glucagon receptor antagonists had reached phase 2

trials and have exhibited robust haemoglobin A1c-reducing efficiency [279], yet the

studies were discontinued due to an elevation in LDL cholesterol, presumably

reflecting the requirement for hepatic glucagon signalling mentioned above.

Aims of the thesis

The role of SLC30A8 in the α-cell has not been characterized. The aim here was to

investigate the effect ZnT8 has on glucose homeostasis and glucagon secretion. In

order to do this mice were generated with α-cell specific deletion or overexpression

of ZnT8.

Page 58: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

58

Chapter 2

Materials & methods

Page 59: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

59

2.1 In vivo animal work

2.1.1 Mice maintenance and generation

Animals were kept in a pathogen-free facility under a 12 h light-dark cycle with

access to water and a standard mouse diet (Lillico Biotechnology). After weaning at

3-4 weeks of age they were housed two to five per separately ventilated cage. The

transgenic mouse strains were maintained on a C57BL/6 genetic background. All

procedures were performed in accordance with UK Home Office regulations under

the Animals (Scientific Procedures) Act 1986 (PPL 70/7349, Dr Isabelle Leclerc).

2.1.2 Generation of transgenic mice

Mice bearing SLC30A8 alleles where exon 1 was flanked by LoxP sites were obtained

from GenOway (Lyon, France). To generate mice with ZnT8 deleted specifically in the

α-cell ZnT8fl/fl mice [272] were crossed to mice bearing the Cre transgene under the

control of a ~0.6 kb fragment of the pre-proglucagon promoter (PPGCre) [280].

2.1.3 Glucose tolerance tests

Mice were fasted overnight before intraperitoneal injection with glucose (1g/kg).

Blood samples were subsequently taken from the tail vein at 0, 15, 30, 60 and 120

min and glucose concentration was measured with an automatic glucometer as

previously described [222].

2.1.5 Insulin tolerance tests

Mice were fasted for 5 h before intraperitoneal injection with insulin (0.5U/kg).

Blood samples were subsequently taken from the tail vein at 0, 15, 30 and 60 min

and glucose concentration was measured with an automatic glucometer as above.

Additionally, plasma (~100 μl) was taken at time 0 and 15 min for glucagon

measurements using a radioimmunoassay (Millipore).

Page 60: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

60

2.1.6 Fasting plasma glucagon measurements

Mice were fasted overnight before taking blood samples of about 40 µl from the tail

vein. Plasma was separated from the cellular components by centrifugation at 12000

rpm for 10 min at 4 °C. A glucagon RIA kit (Millipore) was used to measure glucagon.

2.1.7 Streptozotocin treatments

Mice at the age of 6-7 weeks were treated with 50 mg/kg streptozotocin in a 0.1M

sodium citrate buffer pH 4.5 for five consecutive days following an overnight fast.

Fasting hyperglycaemia (14-19 mM) was evident about 4 weeks later and IPGTT were

performed at this point.

2.1.8 Hypoglycaemic clamps

The clamps were performed by S. Migrenne, C. Magnan (University Paris Diderot-

Paris 7-Unit of Functional and Adaptive Biology (BFA) EAC 7059 CNRS, France) as

described [281].

2.2 In vitro experiments

2.2.1 Islet preparation

Animals were terminated in a CO2 chamber, and death was confirmed by cervical

dislocation. The abdomen was opened along the sternum for extraction of the

pancreas. Isolated pancreata, inflated with collagenase solution at 1 mg/ml (Serva),

were placed in a water bath at 37 °C for 10 min. Cold (~4°C) additive-free RPMI

(Sigma) was added followed by centrifugation at 1000 rpm for 1 min. The

supernatant was removed before adding about 15 ml of cold additive-free RPMI and

the pancreas was resuspended in the media. This pellet was washed three more

times. The islets were then resuspended in 3 ml of Histopaque 119 (Sigma). A

sucrose gradient was then made by adding dropwise 3 ml of Histopaque 1083 and

1077 (Sigma). Finally, RPMI was added dropwise up to 12 ml. The gradient was

Page 61: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

61

centrifuged at 2500 rpm for 20 min. The islets found just below the RPMI layer were

collected and centrifuged at 1000 rpm for 1 min after addition of up to 12 ml RPMI.

The supernatant was removed; islets were resuspended and placed in a Petri dish

containing 15 ml complete RPMI media. After a few hours of recovery at 37 °C islets

were handpicked and placed in 15 ml of fresh media.

2.2.2 Islet dissociation

After overnight culture islets were hand-picked and washed once in 5 ml phosphate

buffered saline (PBS). Then 3-4 ml of Hank’s dissociation buffer (Invitrogen)

supplemented with 0.01% BSA, was added and the islets were centrifuged at 2000

rpm for 2 min. This was repeated once more. The islets were then dissociated using a

P200 pipette in 150 µl of Hank’s buffer and 20 µl 0.1% trypsin solution (Sigma).

Following dissociation, 20 µl of fetal bovine serum (FBS) was added and the cells

were washed with PBS and either resuspended for plating on coverslips coated with

polylysine in 30-40 µl media per coverslip or were incubated with DAPI for 10 min to

stain for dead cells prior to FACS.

2.2.3 Glucagon secretion

Experiments were performed in 12 well plates. Size-matched islets (18) were placed

in 1 ml of 10 mM glucose solution for 30 min in a 37 °C water bath with continuous

and gentle agitation at ~70 revolutions per min. The buffer was then replaced with

fresh KRBH containing either 1 mM or 17 mM glucose for further incubation at 37 °C

for 1 h. The supernatant was collected (secreted glucagon) and the islets were lysed

by sonication for about 10 s in acidified ethanol (500 μl; total glucagon). Secreted

and total glucagon were measured by radioimmunoassay (Millipore). Samples (100

μl) were used for the assay. Sample was mixed with anti-glucagon antibody at 4 °C

for 20-22 h before addition of I125 glucagon for 22-24 h again at 4 °C. Precipitating

agent (Goat anti-Guinea Pig IgG Serum, 3% PEG and 0.05% Triton X-100 in 0.05M

Phosphosaline, 0.025M EDTA, 0.08% Sodium Azide) was then added for 20 min at 4

Page 62: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

62

°C and centrifuged at 4000 rpm for 20 min. The standard curve, total and non-

specific binding values were used to calculate the glucagon content of secreted and

total samples. Secreted glucagon was expressed as a percentage of total glucagon.

2.2.4 Ca2+ imaging with Fluo-2

Islets were incubated (37 °C 95% O2/5% CO2) for 45 min in 10 μM Fluo-2-AM

(Invitrogen) dissolved in DMSO (0.01% wt/vol) and pluronic acid (0.001% wt/vol) in a

bicarbonate-based buffer (120 mM NaCl, 4.8 mM KCl, 1.25 mM NaH2PO4, 24 mM

NaHCO3, 2.5 mM CaCl2, 1.2 mM MgCl2) pH 7.4 containing 3 mM glucose. Following

incubation, 3-4 islets were placed in a perfusion chamber, mounted on a Zeiss

Axiovert confocal microscope and perfused continuously at 37 °C with buffer

containing 1mM glucose followed by 11 mM glucose. Fluo-2 was excited with a 491

nm laser and emitted light was filtered at 525/50 nm. VolocityTM was used for data

capture and ImageJ [282] for image analysis. Traces are shown as normalized

intensity over time (F/Fmin).

2.2.5 Zn2+ imaging with FluoZin-3

Islets were incubated (37 °C 95% O2/5% CO2) for 1 h in 1 μM FluoZin-3-AM

(Invitrogen) dissolved in dimethyl sulfoxide (DMSO) (0.01% wt/vol) in a bicarbonate

buffer containing 11 mM glucose. Islets were transferred into a perfusion chamber

at 37 °C and continuously perifused with buffer. Imaging was done at baseline and

after addition of the Zn2+ chelator N,N,N′,N′-tetrakis(2-pyridylmethyl)

ethylenediamine (TPEN) using the Zeiss Axiovert confocal microscope. FluoZin-3 was

excited with a 491 nm laser and emission monitored at 525/550 nm. Results were

presented as F1/Fmin where F1 was mean baseline fluorescence and Fmin minimum

fluorescence in the presence of the Zn2+ chelator. Traces were analysed using an

Image J macro.

Page 63: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

63

2.2.6 Granular Zn2+ imaging with Zinpyr-4

Islets were incubated as above for 30 min in 10 μM Zinpyr-4 (Santa Cruz) [283].

Imaging was performed on a Nikon Ti Eclipse microscope equipped with an X-Light

spinning disk and Hamamatsu ORCA-Flash4.0 CMOS detector. Static images were

captured at 37 °C using a 63x objective, keeping the exposure time and laser

intensity constant throughout. Zinpyr-4 and RFP were excited using 488 nm and 561

nm lasers, respectively, and signals monitored at 535/50 nm (Zinpyr-4) and 630/75

nm (RFP). Differences in Zinpyr-4 intensity between WT and KO animals was

measured using the corrected total cell fluorescence (CTCF) (integrated density –

(cell area x mean background fluorescence)), as described [284].

2.2.7 Mammalian cell culture

αTC1.9 cells, a mouse glucagonoma cell line provided by the American Tissue Culture

Collection (ATCC) [285], were grown in Dulbecco’s modified Eagle’s medium (DMEM)

without glucose (Sigma) supplemented with 10% FBS, 15 mM Hepes, 0.1 mM non-

essential amino acids, 0.02% BSA, 1.5 g/L sodium bicarbonate, 2 g/L glucose and 1%

penicillin/streptomycin in a 37 °C incubator with 95% O2/5% CO2.

INS-1 (832/13) cells [286], a rat insulinoma cell line, were grown in RPMI medium

(Sigma) containing 11mM glucose, supplemented with 10% FBS, 2 mM glutamine,

1% penicillin/streptomycin, 10 mM Hepes, 1 mM sodium pyruvate and 50 µM β-

mercaptoethanol in a 37 °C incubator with 95% O2/5% CO2.

HEK 293 cells [287], a human embryonic kidney line, were grown in DMEM

containing 25 mM glucose, supplemented with 10% FBS, 1% penicillin/streptomycin

in a 37 °C incubator with 95% O2/5% CO2.

Page 64: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

64

2.2.8 Zn2+ imaging with eCALWY-4: cells

αTC1.9 cells were plated on 24 mm coverslips and cultured overnight to grow. The

following day they were infected with the eCALWY-4 adenovirus, a Förster

resonance energy transfer (FRET) based sensor [288], at a multiplicity of infection of

25-30 for imaging the next day. Cells on coverslips were washed twice in Krebs

Hepes-bicarbonate (140 mM NaCl, 3.6 mM KCl, 0.5 mM NaH2PO4, 0.2 mM MgSO4,

1.5 mM CaCl2, 10 mM Hepes, 2 mM NaHCO3) buffer pre-conditioned with 95:5

O2:CO2 at a pH of 7.4. Solutions of TPEN in ethanol and pyrithione in DMSO were

prepared on the day of the experiment and used at a final concentration of 25 mM

and 1 mM respectively. Cells were maintained at 37 °C using a heating stage and

perifused with Krebs Hepes buffer initially with no additions. Calibration of the FRET

probe was performed by perfusing with the Zn2+ chelator TPEN followed by a

combination of Zn2+ (100 µM) and the Zn2+ ionophore pyrithione. The acquisition of

images was done with a 100 ms exposure time and 433 nm excitation. The

microscope was fitted with a monochromator and for FRET measurements a dichroic

mirror and two emission filters had to be used. Cells were imaged with a 40x oil

immersion objective.

The FRET-based probe undergoes a large ratiometric change following Zn2+ binding

[288]. The probe has a donor and an acceptor molecule and a conformational change

in the sensor domain results in a change in the ratiometric fluorescence signal from

these molecules. Dissociation of the two molecules means the donor emission is

detected upon excitation while when the two are in proximity the acceptor emission

is the one primarily detected because of FRET from the donor to the acceptor. In

eCALWY-4 the metal binding domains are Atox-1 and WD-4 linked to the

fluorophores Cerulean and Citrine respectively [288]. The metal binding domains are

linked by an amino acid linker. At low Zn2+ levels the metal binding domains are

separated from each other so that the two fluorophores can be close to each other

for generation of energy transfer. When Zn2+ levels increase, the metal binding

Page 65: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

65

domains are in closer proximity separating the two fluorophores thus reducing

energy transfer (Figure 2.1).

Figure 2.1 – eCALWY sensor design. Atox1 and WD4 were used as metal binding domains.

Changes in conformation upon Zn2+ binding cause in energy transfer and thus fluorescence

that can be used to determine Zn2+ levels. From Vinkenborg et al [288].

2.4.4 Amplification and purification of adenoviral eCALWY4

AD293 cells (ATCC) were infected with 1 μl of the virus. After three rounds of

amplification the cells were harvested and were subjected to a freeze-thaw cycle to

release the virus. The cells were centrifuged at 5000 rpm at 4°C and the supernatant

was collected. Saturated CsCl (5.4 ml) in PBS was added to 9 ml of supernatant in a

13.5 Sorvall ultracentrifuge tube and was centrifuged for 10-16h at 4 °C at 65 000

rpm. The viral band was removed with a 23G needle, in dialysis tubing and placed in

500 ml dialysis buffer at 4 °C. The buffer was changed every hour 4 times. The virus

was then stored in storage buffer in 1:2 dilution in 40 μl aliquots at -80 °C.

2.4.5 Virus titration

AD 293 cells were plated ~24 h prior to infection on a 24-well plate. The cells were

infected with the eCALWY-4 adenovirus in serial dilutions when they reached ~70 %

confluency. After 24 h infection the cells were imaged with a 20x objective using an

Page 66: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

66

Olympus epifluorescence microscope, counting the total number of green cells per

well which was used to calculate the MOI.

2.2.9 Immunostaining: cells

Cells were plated onto coverslips and the following day culture media was removed

before washing three times with PBS. Cells were fixed in 4% paraformaldehyde (PFA)

(w/v) in PBS for 15 min at room temperature and then washed with PBS. Cells were

permeabilized with 0.1% Triton X-100 in PBS for 10 min and subsequently washed

with PBS. 1% BSA in PBS was used as a blocking solution for 30-60 min at room

temperature. Incubations with primary antibodies at the appropriate dilutions were

performed at 4 °C overnight or for 48 h using 1% BSA in PBS. After more washes in

PBS cells were incubated with secondary antibodies at a 1:1000 dilution for 1 h at

room temperature on a shaker (30 revolutions per min) Coverslips were mounted on

microscope slides (Thermo Scientific) using DAPI containing mounting medium

(Vector Labs).

2.2.10 Immunostaining: islets

Fixation of islets was done in 4% paraformaldehyde (PFA) for two hours at room

temperature or overnight at 4 °C. After washing with PBS islets were incubated with

primary antibodies in 5% BSA plus 0.1% Triton X-100 in PBS overnight or for 48 h at

4°C. After more washes in PBS, islets were incubated with secondary antibodies at a

1:1000 dilution for 1h at room temperature on a shaker. Islets were transferred onto

slides before adding a drop of mounting medium (Vectashied with DAPI) and a

coverslip on top. The islets were imaged with a 20x magnification objective using a

confocal microscope.

Page 67: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

67

2.2.11 Immunostaining: slices

Freshly isolated pancreata were fixed in 10% formalin (w/v) overnight at 4 °C,

embedded in wax and sliced into 5 µm sections. They were then dewaxed for 20 min

in Histoclear (Sigma) and rehydrated by washing for 5 min in 100, 95 and 70%

ethanol and distilled water. The sections were treated with vectors antigen solution

(Vectors Labs) 3-4 times washing with distilled water in between treatments. They

were then washed twice in PBS and blocked in PBS containing 0.1% Triton X-100, 2%

BSA and 2% goat and donkey serum for 60 min at room temperature. Primary

antibodies (anti-guinea pig insulin 1:200 (DAKO), anti-rabbit glucagon 1:100 (Sigma))

in PBS were then applied on the slides for overnight incubation at 4°C in a moist box.

The next day, slides were washed 3 times in PBS before incubation with secondary

antibodies (488 goat anti-guinea pig 1:1000, 568 donkey anti-rabbit 1:500) in PBS at

room temperature. There were some final washes in PBS before adding a drop of

Vectashield with DAPI on the slide and covering with a rectangular coverslip. The

slides were flat at 4°C and allowed to dry. Imaging was done on a Zeiss AxioObserver

widefield using a 40x objective. The α:β cell ratio was determined by subjecting

every islet image to manual thresholding (Image J macro written by Dr Pauline

Chabosseau) and dividing the α- cell area by the β-cell area.

2.2.12 Quantification of pancreatic glucagon content

Pancreata were extracted from mice fasted for 5 h, split into two, weighed and

homogenized using a Polytron homogenizer in 5 mL acidic ethanol (1.5% HCl, 70%

ethanol). They were incubated overnight at 4°C and homogenized further by

sonication. Another overnight incubation followed and the samples were centrifuged

at 2000 rpm at 4 °C for 15 min. The solution was removed and neutralized with an

equal volume of 1 M Tris-HCl. The samples were then diluted and glucagon

concentration was measured using a Homogeneous Time Resolved Fluorescence

(HTRF) glucagon assay (Cisbio) in a PHERAstar reader. The total amount of glucagon

was then calculated as ng/mg of pancreas.

Page 68: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

68

2.2.13 Electron microscopy

Islets (~100-150) were handpicked, washed in PBS and centrifuged at 500 rpm for 30

s in a 1.5 ml eppendorf. PBS was removed and the islets were fixed in 500 μl freshly

prepared Vincenzo’s fixative (2% PFA, 2.5% glutaraldehyde, 3 mM CaCl2, 0.1 M

sodium cacodylate (pH 7.4)) for 30 min at 37 °C. The islets were centrifuged again

and the fixative was replaced with more fixative for 2 h at room temperature. Finally

the fixative was replaced again with fresh fixative and left overnight at 4 °C before

being resuspended in PBS.

2.3 Molecular biology

2.3.1 Genotyping

Ear clips were taken from mice using an ear puncher. DNA was extracted by addition

of 75 µl of alkaline lysis reagent (25 mM NaOH, 0.2 mM EDTA pH 8.0) to the ear

punches and incubating at 98 °C for 1 h. After this, neutralisation reagent (75 µl; 40

mM Tris HCl) was added prior to centrifugation at 14000 rpm 1 min. Extracted DNA

was ready to use in genotyping PCRs.

2.3.2 RNA extraction

Cells or islets were washed once in PBS followed by addition of 1ml TRIzol

(Invitrogen) for cells and 500 µl for islets in a 1.5 ml Eppendorf tube. After this 200

µl of chloroform (Sigma) per ml of TRIzol were added, the tube was vortexed and left

at room temp for 2-3 min until two layers formed. Samples were centrifuged at

12000 rpm for 15 min at 4 °C. The upper aqueous phase was removed and RNA was

precipitated by adding 400 µl isopropanol (Sigma) per ml of TRIzol used initially to

the retrieved aqueous phase. The samples were vortexed and incubated at room

temperature for 15 min before spinning for 15 min at 4 °C. The pellet was washed

twice in 1 ml 75% ethanol and centrifuged at 7500 rpm for 5 min. The dried pellet

Page 69: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

69

was finally resuspended in nuclease-free water. The same quantity of RNA was used

from each sample to perform an RT-PCR and cDNA synthesis using random primers.

2.3.3 cDNA conversion

cDNA was generated from RNA using a High Capacity Reverse Transcription kit

(Applied Biosystem). A 1X reaction included 2 μl of each 10X RT Buffer, 10X RT

Random Primers, 0.8 μl 25X dNTP Mix (100mM), 1 μl Multiscribe Reverse

Transcriptase. The amount of H20 depended on the amount of RNA used. The

reaction was made up to 20 μl with the addition of nuclease-free H20. A maximum of

1μg RNA was converted to cDNA. For conversion the following protocol was carried

out in a thermocycler: 10 min 25 °C, 2 h 37 °C, 5 min 85 °C, hold 4 °C.

2.3.4 Real Time PCR

For gene expression measurements cDNA (2μl) from the RT-PCR reaction was used

as template for quantitative real time PCR using a Fast SYBR Green Master Mix

(Invitrogen). A 1X reaction was made up of 2 μl template, 6 μl SYBR Green, 0.35 μl

forward and 0.35 μl reverse primers, 3.3 μl H20. Each sample was amplified in

duplicates. The reaction was initiated at 50 °C for 2 min followed by the activation

and pre-denaturation step at 95 °C for 10 min. The run was made up of 40 cycles of

15 s at 95 °C and 1 min at 60 °C. mRNA levels were expressed as ΔCt and β-actin was

used as an endogenous control for normalisation.

2.3.5 Fluorescence activated cell sorting (FACS)

After overnight incubation, 250-300 islets of the same genotype were handpicked in

a 15 ml falcon and washed twice with 5 ml Hank’s based cell dissociation buffer

(Invitrogen) containing 0.1% BSA. They were then centrifuged at 2000 rpm for 2 min.

Islets were dissociated by repeated pipetting (80-100 times) in 150 μl of the solution

and 20 μl trypsin. The reaction was stopped with the addition of 20 μl FBS (Seralab).

The cells were washed in 1ml PBS, centrifuged at 2000 rpm for 1 min and

Page 70: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

70

resuspended in 500 μl PBS and 1% FBS. 1 µg/ml DAPI (Roche) was added to stain for

dead cells. After a 10 min incubation the cells were filtered through a 35 μm cell

strainer. RFP+ and RFP- cells were recovered in 500 μl TRIzol after sorting on a BD

FACS ARIA III (BD Bioscience).

2.3.6 FACS analysis

Islets were dissociated into single cells as previously described, washed in PBS and

centrifuged at 600 g for 2 min. Cells were incubated in 50 μl 1:1000 near IR dead cell

stain (Life technologies) for 20 min at 4 °C, washed with PBA (PBS, 1% BSA, 0.1%

azide) and fixed in 2% PFA for 10 min at room temperature. They were then washed

twice with PBA and once with Saponin (0.025 % in PBA) before 10 min incubation

with Saponin at room temperature. Cells were incubated with primary antibodies

against ZnT8 (Mellitech), insulin (DAKO) and glucagon (Sigma) in 1:100 dilution in

Saponin for 20 min at room temperature. They were then washed twice in Saponin

before incubation with secondary antibodies (anti-mouse AF 405, anti-guinea pig AF

488, anti-rabbit AF 640) for 20 min at RT. Next were two final washes in Saponin

before resuspending in PBA. The samples were run on a BD Fortessa Flow

Cytometer (BD Bioscience).

2.4 Cloning and virus production

2.4.1 Competent bacteria

A trace of competent XL-1 Blue bacteria was taken with a sterile inoculating loop,

streaked out on a luria broth (LB) agar plate and incubated at 37 °C overnight. A

single colony was picked to inoculate 10 ml LB and incubated in a shaker at 37 °C

overnight. Overnight culture of 1 ml was added to pre-warmed LB and incubated on

the shaker at 37 °C until an appropriate density was reached (approximately 3.5 h).

The culture was cooled down on ice for 5 min before transfer into 50 ml centrifuge

tubes and centrifugation at 4000 g for 5 min at 4 °C. The supernatant was discarded

Page 71: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

71

and cells were kept on ice. The following steps were performed in the cold room.

Cells were resuspended in 30 ml ice-cold TFB1 buffer (100 mM RbCl, 50 mM MnCl2,

30 mM KC2H4O2, 10 mM CaCl2, 15% glycerol) and kept on ice for 90 min. Cells were

collected by centrifugation at 4000 g for 5 min at 4 °C and then resuspended in 16 ml

ice cold TFB2 buffer (10 mM MOPS, 10 mM RbCl, 75 mM CaCl2, 15 % glycerol).

Aliquots (200 µl) were prepared in 1.5 ml Eppendorf tubes. These were frozen down

in a dry-ice ethanol bath. Competent cells were stored at -80 °C.

2.4.2 Transfection

Cells were seeded in 6-well plates and transfected the following day using

Lipofectamine-2000 (Invitrogen). Plasmid DNA (1 µg) and 3 µl Lipofectamine were

used per well. These were mixed separately into Optimem (100μl) and incubated at

room temperature for 5 min before mixing and further incubation for 20 min. 200 µl

of this solution and 1 ml Optimen were added to each well before culture overnight

at 37 °C in a humidified incubator. This was then replaced with normal culture

media. Imaging was done 48 h following transfection.

2.4.3 Bacterial transformation

Plasmid DNA (~10 ng) was added to 100µl of competent cells. The bacteria were left

on ice for 15 min before heat shock at 42°C for 1 min. Next, the cells were incubated

for 1h at 37 °C with constant shaking after addition of 300 µl of super optimal broth.

Cells (50μl) were plated on a warm LB/agar plate which included the suitable

antibiotic at a concentration of 100µg/ml.

2.5 Statistical analysis

Student’s t-tests were used to determine any significant difference between two

independent variables. Interactions between multiple treatments were examined

using two-way ANOVA (with adjustment for repeated measurements as necessary)

Page 72: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

72

prior to pairwise comparisons with Bonferroni’s post hoc test. Analysis was

performed with GraphPad Prism 6.0. A value of P<0.05 was considered significant.

Data are expressed as means ± SEM.

Page 73: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

73

2.6 Primers

Gene

Sequence

ZnT1 FOR: CAGGCAGAGCCAGAAAAATTG

REV: TGGATGAGATTCCCATTTACTTGTAC

ZnT2 FOR: CCCGACCAGCCACCAA

REV: CCAAGGATCTCGGCTCGAT

ZnT3 FOR: GGAGGTGGTTGGTGGGTATTT

REV: CAAGTGGGCGGCATCAGT

ZnT4 FOR: CATCGCTGCCGTCCTCTAC

REV: ATTTGCCATGTATCCACCTACAAG

ZnT5 FOR: ACTGTTTGCTGCCCTGATGAG

REV: CTCTATTCGGCCATACCCATAGG

ZnT6 FOR: ACGTCTCTGAAGCTGCTAGTACGA

REV: CCACACAAGCTGCGGCTAA

ZnT7 FOR: TGTGCCTGAACCTCTCTTTCG

REV: GCCTAGGCAGTTGCTCCAGAT

ZnT8 FOR: TGCACAGTCTACACATCTGGTCACT

REV: TGGCTGGCAGCTGTAGCA

ZnT9 FOR: GCACTGGGCATCAGCAAAT

REV: GAAAAGCCGTACGGGTGAGA

MT1 FOR: CCTTCTCCTCACTTACTCCGTAGC

REV: GGAGCCGCCGGTGGA

MT2 FOR: TCCTGTGCCTCCGATGGAT

REV: TGCAGGAAGTACATTTGCATTGT

MT3 FOR: AAGTGCAAGGGCTGCAAATG

REV: CACAGTCCTTGGCACACTTCTC

Actin FOR: CGAGTCGCGTCCACCC

REV: CATCCATGGCGAACTGGTG

Page 74: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

74

Glucagon FOR: CCAAGAGGAACCGGAACAAC

REV: CCTTCAGCATGCCTCTCAAAT

Insulin FOR: CGTGGCTTCTTCTACACACCC

REV: AGCTCCAGTTGTGCCACTTGT

Abcc8-SUR1 FOR: GCCTACGCATCTCAGAAACCA

REV: CCATCTTGTACCTTTGCTTATTGAAG

Arx FOR: TGGGTCTGAGCACTTTTCTAGGA

REV: GAAAAGAGCCTGCCAAATGC

Gck FOR: CACAATGATCTCCTGCTACTATGAAGA

REV: AGCCGGTGCCCACAATC

Pcsk2 FOR: GAGTCCGAAAGCTCCCCTTT

REV: TTTGCAAGCCCATTGTGATAAA

Human ZnT8 FOR: CTGTCATCGAAGCCTCCCTC

REV: AAGGGCATGCACAAAAGCAG

Cre FOR: ATGTCCAATTTACTGACCG

REV: CGCCGCATAACCAGTGAAAC

ZnT8 flox FOR: AGTTATTGACTGAACACACCTATCTTA

REV: GCTATATACTCTTCCACTCAGCTACAT

RFP FOR: CTACAGGAACAGGTGGTGG

REV: CTGTTCCTGGGGCATGGC

Beta catenin FOR: AAGGTAGAGTGATGAAAGTTGTT

REV: CACCATGTCCTCTGTCTATTC

GlurtTA FOR: CATAAACGGCGCTCTGGAATTACTCAATGGAGTCG

REV: GCCGAGATGCACTTTAGCCCCGTCGCGATGTGAGA

Luciferase FOR: CAACTGCATAAGGCTATGAAGAGA

REV: ATTTGTATTCAGCCCATATCGTTT

Page 75: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

75

2.7 Antibodies

Primary

Antibody Species Company Dilution

ZnT8 Rabbit Mellitech 1:200

Glucagon Mouse Sigma 1:1000

Glucagon Rabbit Santa Cruz 1:200

Insulin Guinea pig Santa Cruz 1:200

RFP Rabbit Clontech 1:100

Secondary

Antibody Species Company Dilution

Anti-mouse Alexa 568 Goat Invitrogen 1:1000

Anti-rabbit Alexa 488 Goat Invitrogen 1:1000

Anti-guinea pig Alexa

488 Donkey Invitrogen 1:1000

Anti-mouse Alexa 488 Donkey Invitrogen 1:1000

Anti-rabbit 568 Goat Invitrogen 1:1000

Page 76: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

76

Chapter 3

ZnT8 deletion from a subpopulation of α-cells enhances

hypoglycaemia-induced glucagon secretion

Page 77: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

77

3.1 Introduction

When the function of the hormone-producing α- and/or β-cells becomes impaired

this is likely to impact the regulation of glycaemia and could eventually cause

disease. These changes are associated with both T1D and T2D (see Chapter 1).

In addition to alterations in insulin release [289], glucagon release is also defective

and contributes to diabetic hyperglycaemia and impaired glucose homeostasis in

advanced stages of both disease types. In both cases, the α-cell no longer responds

normally to glucose concentrations so that the counter-regulatory response to

hypoglycaemia is weakened. Conversely, raised glucose levels, which should inhibit

glucagon secretion, elicit inappropriately high glucagon responses [290].

Genome-wide association studies examine SNPs across the genome and seek to

assess which, if any, may contribute to disease risk. Such research has so far led to

the discovery of ~90 loci in the human genome which appear to have a significant

impact on T2D risk [203, 205, 212, 291, 292]. Few of the SNPs have a known

functional impact on the amino acid sequence, with the majority of them located in

non-coding parts of the genome.

One of these loci harbours SLC30A8 which encodes the Zn2+ efflux transporter ZnT8.

The expression of this gene is largely restricted to the pancreatic islets where the

transporter resides on the membrane of the dense core secretory granule in α- and

β-cells [203].

A non-synonymous SNP rs13266634 results in an amino acid substitution from a

tryptophan to an arginine in the cytoplasmic C-terminus of the protein. This risk

variant has one of the strongest effects on T2D risk, 15% per inherited allele. The

transcript is believed to be translated into a protein with lower Zn2+ activity making

the transporter less efficient at Zn2+ accumulation across the membrane into the

granule [222, 259]

Page 78: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

78

A considerable amount of research has been focused on endeavouring to take

information from GWAS towards eventual clinical exploitation by exploring the

function of the associated genes in vivo in mouse models. Various groups have

looked into the effect of modulating ZnT8 expression in the mouse by deleting the

gene either globally or specifically in the β-cell. Given ZnT8’s subcellular localization,

and the importance of Zn2+ in insulin biosynthesis and storage [267], the role of the

transporter has largely, if not almost completely, been studied in the context of the

β-cell and insulin secretion.

3.2 Aims

Although many research groups have done extensive work looking into the role of

ZnT8 in the β-cell, work on the α-cell is lacking and calls for further investigation.

Here we aimed to generate a mouse deleted selectively for ZnT8 in the α-cell and

investigate how this affects glucose tolerance and glucagon secretion in vivo.

Page 79: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

79

3.3 Results

3.3.1 Expression of ZnT8 in the mouse α-cell

Dissociated islets from C57BL/6 mice were stained for glucagon and ZnT8 and

imaged by spinning disk confocal microscopy. ZnT8 immunoreactivity was present in

glucagon-positive cells and there was colocalization between the two stains.

However, the degree of colocalization was far from complete (Pearson correlation=

0.79) (Figure 3.1A). The presence of ZnT8 in the α-cell was also confirmed at the

mRNA level after extracting total RNA from fluorescence activated cell sorting (FACS)

purified α-cells. This revealed that ZnT8 is the most highly expressed member of the

ZnT family of transporters in these cells (Figure 3.1B)

Figure 3.1 – ZnT8 expression in the α-cell. (A) Dissociated islets cells were stained for

glucagon using anti-glucagon (Sigma, 1:1000, red) and anti-ZnT8 (Mellitech, 1:200, green) to

confirm the presence of ZnT8 at the protein level. Scale bar = 10 µm (B) Total RNA was

extracted from FACS-sorted cells to verify the expression of ZnT8 in the α-cell at the mRNA

level. Data from n=3 mice

Zn T1 Zn T2 Zn T3 Zn T4 Zn T5 Zn T6 Zn T7 Zn T8 Zn T9

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

-a

cti

n)

Glucagon ZnT8 A

B

Page 80: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

80

3.3.2 Generation, breeding and phenotyping of mice null for ZnT8 in the α-cell

Animals bearing floxed ZnT8 alleles, originally obtained from GenOway (Lyon,

France) and backcrossed with C57BL/6 mice, were maintained on a standard chow

diet. The strategy followed by GenOway to generate ZnT8fl mice required the

insertion of LoxP sites, one upstream of exon 1 and a second one connected to a FRT

flanked neomycin selection cassette upstream of exon 1 and within intron 1 of the

Slc30a8 gene. The construct was homologously recombined in SV129-derived

embryonic stem cells at the ZnT8 site and was then injected into C57BL/6

blastocytes, before being transferred into pseudopregnant C57BL6 females. Crossing

with a flp-expressing strain then removed the neomycin cassette. The resulting

animals were further backcrossed 6 times to C57BL/6 mice.

The ZnT8fl/fl mice were crossed with mice carrying a Cre transgene under a 0.6 kB

fragment of the pre-proglucagon promoter (PPG) [280]. Further crossing was then

needed to generate homozygous mice for the floxed gene containing a single copy of

the transgene, rendering them ZnT8 KO (Figure 3.2A). The genotypes of the offspring

were identified by PCR analysis of genomic DNA extracted from ear biopsies. The

samples were tested for the presence of ZnT8 floxed alleles (Figure 3.2B) and Cre

transgene to determine the KO animals and those that would act as control (Figure

3.2C). This strain of mice was used in in vivo experiments and glucagon secretion

assays.

Morphologically α-cells are almost identical to the more abundant β-cells. The lack of

reliable and precise methods of distinguishing α-cells in the intact islet makes

research on α-cells challenging. Thus, this project also utilized transgenic mice

bearing a fluorescent protein for identification of α-cells in situ generating a triple

transgenic. ZnT8fl/fl mice with the Cre transgene under the control of the PPG

promoter were crossed with mice containing a stop-flox red fluorescent protein

variant (tandem dimer RFP, tdRFP/RFP) at the Rosa26 locus. In the presence of Cre

recombinase, recombination events flip the RFP construct from the antisense

Heterozygote

Page 81: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

81

orientation into the sense. Another recombination event then takes place during

which the stop-flox cassette is removed allowing transcription driven from a CAG

promoter to read through into the properly-oriented tdRFP gene [293]. Using a Cre

maintained under the control of the PPG promoter prevented expression of RFP in

any other tissue or cell type other than α-cell [280]. The breeding strategy followed

can be seen in Figure 3.2B. In addition to this colony a separate one had to be bred

with mice WT for the ZnT8 locus but still bearing the PPGCre transgene and RFP for

use in in vitro experiments as controls (Figure 3.2C).

Figure 3.2 – Generation of αZnT8 KO mice and genotyping by PCR. (A) Schematic

representing the recombination event taking place to generate αZnT8 null mice in the

presence of Cre. (B) PCR genotyping for the floxed or WT allele. The band at ~ 330 bp is from

the WT allele while the larger band at ~ 460 bp is from the floxed allele of ZnT8. (C) PCR

genotyping for the Cre transgene to identify WT and KO animals. The band at ~ 330 bp

verifies the presence of Cre and the band at ~220 bp is that for β-catenin used as PCR

control.

200 bp 300 bp

Control KO

ZnT8 targeted allele

LoxP LoxP

Exon1 Exon2

LoxP

Exon2

Cre

A

B C

300 bp 500 bp

Homozygote

WT

Heterozygote

Page 82: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

82

A C B

Figure 3.3 – Breeding strategy followed to generate αZnT8 KO mice and necessary controls

(A) for in vivo experiments that required littermate controls but not RFP, (B) for RFP-tagged

ZnT8 KO α-cells, (C) for RFP tagged α-cells with WT ZnT8 alleles.

3.3.3 Recombination efficiency and ZnT8 deletion

To assess the degree of recombination occurring in mice containing the Cre

transgene, initially whole islets from Cre+RFP+ mice were stained for RFP alongside

glucagon or insulin (Figure 3.4A). Imaging of these islets showed that only about 45%

of glucagon-positive cells expressed RFP. An almost negligible (2%) proportion of

insulin-positive cells was also labelled with RFP but this was most likely due to non-

specific background staining (Figure 3.4B). Additionally, islets positive for RFP from

WT and KO islets were stained for ZnT8 and glucagon or insulin to examine the

deletion of ZnT8 and not just the recombination efficiency. Immunohistochemical

analysis of the latter islets showed that ZnT8 immunoreactivity (Figure 3.5A) was

selectively eliminated from KO and not WT mice. However, only about half the

population of RFP+ cells showed a reduction in ZnT8 (Figure 3.5B). Insulin-positive

cells did not show a reduction.

Page 83: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

83

W T K O

0

5 0

1 0 0

Pe

rce

nta

ge

(%

)

****

KO

WT

RFP ZnT8 A B

Figure 3.4 – Recombination in α-cells. (A) Representative images of PPGCreRFP+ islets were

stained for RFP (1:100) insulin (1:200) and glucagon (1:1000). (B) Images of stained islets

were used to calculate the percentage of glucagon and insulin positive cells expressing RFP.

Scale bar = 50µm **P<0.01, n=5-8 islets

Figure 3.5– ZnT8 deletion in α-cells. ZnT8 immunoreactivity (1:200) was assessed in WT and

KO islets expressing RFP. (A) Representative images showing ZnT8 in green and endogenous

RFP fluorescence in red. Scale bar = 10µm (B) The bar graph show the proportion of RFP+

cells that were also positive for ZnT8. n= 80-90 cells from 5-6 islets per genotype, Student’s

t-test, ****P<0.0001

To then get a more accurate quantification of the extent of deletion, isolated RFP+

islets from WT and KO mice were dissociated, fixed and stained for insulin, glucagon

and ZnT8 for FACS analysis (Figure 3.6A,B). This showed that approximately 30% of

glucagon-positive cells expressed RFP (Figure 3.6C), reflecting recombination at the

Rosa26 locus where the tdRFP reporter is located. Although clearly recombination in

G lu+

R F P+

In s+

R F P+

0

1 0

2 0

3 0

4 0

5 0

**

Pe

rce

nta

ge

(%

)

A B

Page 84: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

84

these cells did occur at this locus, only 50% of RPF+ cells showed deletion of ZnT8

and thus recombination at the Slc30a8 locus. This suggests that Cre is more efficient

at recombination at the RFP than at the ZnT8 locus.

In addition, ZnT8 was found to be absent from a small percentage of glucagon-

positive RFP- cells indicating that recombination happened in a small proportion of α-

cells not expressing RFP (Figure 3.6D) Concomitantly, insulin-positive cells did not

show any changes in ZnT8 levels and they were all ZnT8-positive (Figure 3.6D).

Overall, ZnT8 expression was eliminated in ~15% α-cells (Figure 3.6D).

Such a low degree of deletion made us look into the use of other promoters as a

means of deleting ZnT8 in a greater proportion of α-cells. In the lab there was an

alternative strain that could have been used, with Cre again under a longer fragment

(>100 kbp) of the PPG promoter (iGluCre), present in a bacterial artificial

chromosome construct [294]. Islets from these iGluCre mice crossed with RFP mice

were extracted and stained to check for the level of recombination in this model. It is

evident from the images of islets stained for RFP and insulin (Figure 3.7) that RFP

expression is not restricted to α-cells since there is considerable double staining of

RFP and insulin making this delete strain unsuitable for use in this study.

3.3.4 Body weight and growth

αZnT8 KO mice were weighed regularly to check whether they had any variances in

weight with respect to control animals. Across all the ages examined, WT and KO

mice had a comparable body weight (Figure 3.8). Additionally, all the mice appeared

healthy with no obvious differences in growth.

Page 85: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

85

Figure 3.6 – FACS analysis of WT and KO islets. Single islets following dissociation of RFP+

WT and KO islets were fixed, permeabilized and stained for insulin, glucagon and ZnT8. Cells

were stained prior fixation with a live/dead stain to only allow the inclusion of cells that

Zn T 8+ - c e lls Zn T 8

+ - c e lls

0

2 0

4 0

6 0

8 0

1 0 0

* * *

% o

f c

ell

s

R F P+

R F P+

Zn T 8+

R F P-

Zn T 8+

0

2 0

4 0

6 0

8 0

1 0 0 W T

K O

* * * *

% o

f c

ell

s

B

A

1) Single cells

2) Live

β-cells – RFP α-cells – RFP RFP+ α-cells – ZnT8

RFP_

α-cells – ZnT8

C D

1) Single cells

2) Live

β-cells – RFP

β-cells – ZnT8

α-cells – RFP

α-cells – ZnT8

RFP+ α-cells – ZnT8

RFP_

α-cells – ZnT8

β-cells – ZnT8 α-cells – ZnT8

Page 86: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

86

were alive just before fixation. Representative traces showing expression analysis of cells

from WT (A) and KO (B) islets. The red traces denote the positive populations while blue

traces the negative populations. (C) Expression of RFP and ZnT8 from WT and KO islets as a

percentage of live cells. (D) ZnT8 expression in glucagon- and insulin-positive cells. Data are

mean ± SEM, n=3, two-way ANOVA, *P<0.01, ***P<0.0005.

3.3.5 Glucose homeostasis

In order to investigate the effect of deleting ZnT8 in α-cells on glucose homeostasis,

and particularly to detect whether these mice may display altered glucose tolerance,

we carried out intraperitoneal glucose tolerance tests (IPGTTs) on αZnT8 KO mice.

Littermates lacking the Cre transgene were used as controls to minimize variability

(Colony 1, Figure 3.3). It has been reported that age differences can have a huge

impact on glucose homeostasis in rodents [295] so IPGTTs were performed at 8, 12

and 16 weeks of age in case any effect might be age dependent. A glucose load of 1

g/kg was injected intraperitoneally and circulating glucose levels monitored at

various time intervals by taking blood from the tail. As can be seen in Figure 3.9

there was no discernible difference at any age examined.

Figure 3.7 – Immunostaining of iGluCre

islets. RFP+ islets bearing Cre under the

alternative longer PPG promoter [294]

were fixed and stained for insulin.

Endogenous RFP can be seen in red and

insulin in green.

insulin merge RFP

Page 87: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

87

0 3 0 6 0 9 0 1 2 0

0

5

1 0

1 5W T

K O

T im e (m in )

Glu

co

se

(m

M)

0 3 0 6 0 9 0 1 2 0

0

5

1 0

1 5

T im e (m in )

Glu

co

se

(m

M)

W T

K O

0 3 0 6 0 9 0 1 2 0

0

5

1 0

1 5

T im e (m in )

Glu

co

se

(m

M)

W T

K O

A C B

Figure 3.8 – Weight measurements. The mice were weighed regularly shortly after they

were weaned up to the age of 16 weeks. There was no difference between WT and KO mice

either in females (A) or males (B).

Figure 3.9 – Glucose homeostasis. Intraperitoneal glucose tolerance tests were performed

on WT and αZnT8 KO mice fed a normal chow diet after an overnight fast at 8 (A) 12 (B) and

16 (C) weeks of age. n=8-15 mice/genotype

Another factor investigated was whether fasting conditions elicited a different

response from the KO mice. Blood glucose levels were measured after an overnight

fast and blood samples were taken from the tail vein for serum glucagon

measurements (Figure 3.10). Again no difference was observed in fasting blood

glucose levels or fasting glucagon compared to WT controls.

A g e (w e e k s )

We

igh

t (g

)

4 8 1 2 1 6

0

1 0

2 0

3 0W T

K O

A g e (w e e k s )

We

igh

t (g

)

4 8 1 2 1 6

0

1 0

2 0

3 0W T

K O

A B

Page 88: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

88

Figure 3.10 – Fasting glucose and glucagon. The mice were fasted overnight and fasting

plasma glucose was measured using a glucometer (A). Plasma was taken for glucagon

measurements (B). n=7-8

3.3.6 Insulin sensitivity

In an effort to impose extra stress on the α-cell, and to test whether ZnT8 absence

affects insulin sensitivity, insulin tolerance tests (ITTs) were performed on mice of 8

weeks of age (Figure 3.11A). Blood glucose was measured at regular intervals and,

similarly to IPGTTs, there were no significant differences between WT and KO mice.

During the ITTs blood was collected at 0 and 15 min for plasma glucagon

measurements. The glucagon levels of WT and KO mice were very close at both time

points (Figure 3.11B).

Figure 3.11 – Insulin sensitivity. (A) Intraperitoneal insulin tolerance tests performed on 8

week old mice fasted for 5 hr and injected with 0.5U/kg insulin. (B) Plasma glucagon

measured at 0 and 15 min. n=13-14, *p<0.05

T im e (m in )

[glu

ca

go

n]

pg

/ml

0 1 5

0

5 0

1 0 0

1 5 0W T

K O*

*

0 1 5 3 0 4 5 6 0

0

5

1 0

1 5

T im e (m in )

Glu

co

se

(m

M)

W T

K O

A B

W T K O

0

2

4

6

8

Glu

co

se

(m

M)

W T K O

0

2 0

4 0

6 0

8 0

1 0 0

Glu

ca

go

n (

pg

/sa

mp

le)

A B

Page 89: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

89

3.3.7 Streptozotocin treatment

In order to assess whether the deletion of ZnT8 might influence glucose homeostasis

in the context of diabetes, ZnT8 WT and KO mice at 6-8 weeks were rendered

diabetic by administration of streptozotocin (STZ), an agent which is selectively toxic

for the β-cell. STZ enters the cell through the GLUT2 transporter which is not

normally expressed in the α-cell [296]. The agent was administered to mice via the

intraperitoneal route for five consecutive days at a dose of 50 mg/kg. Glucose

homeostasis was assessed 4 weeks later by performing an IPGTT after an overnight

fast. The mice displayed fasting hyperglycaemia as expected though there were no

differences in blood glucose levels between WT and KO mice throughout the

duration of the test (Figure 3.12)

Figure 3.12 – Streptozotocin treatments.

Mice were injected with 50mg/kg

streptozotocin daily for 5 consecutive

days following overnight starvation.

IPGTTs were performed on WT and KO

mice 4 weeks later when the mice

showed signs of fasting hyperglycaemia.

n=8-9

3.3.8 Hypoglycaemic clamps

Since direct intraperitoneal insulin injections of insulin (Figure 3.11) could only be

used safely to produce relatively mild hypoglycaemia (4-5 mM) an alternative

approach was used to safely elicit a more marked reduction in blood glucose (2-3

mM) where glucagon secretion is likely to be more strongly stimulated. These

experiments were performed in Prof Christophe Magnan’s laboratory (University

Paris Diderot-CNRS UMR8251). To examine the in vivo response of these mice to low

glucose conditions the mice were rendered hypoglycaemic by infusion of insulin

(Figure 3.13A). Male mice did not exhibit any differences in the rate of glucose they

0 3 0 6 0 9 0 1 2 0

0

1 0

2 0

3 0

4 0W T

K O

T im e (m in )

Glu

co

se

(m

M)

Page 90: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

90

0 9 0

0

2 0 0

4 0 0

6 0 0

8 0 0W T

K O

T im e (m in )

Pla

sm

a g

luc

ag

on

(p

g/m

l)

**

**

0 1 0 2 0 3 0 4 0 5 0 6 0 7 0 8 0 9 0

0

2

4

6

8

1 0

T im e (m in )

Gly

ca

em

ia (

mM

)

W T

K O

2 0 3 0 4 0 5 0 6 0 7 0 8 0 9 0

0

1 0

2 0

3 0

4 0

T im e (m in )

GIR

(m

g/m

in/k

g)

**

*

W T

K O

A

C

B

needed to maintain hypoglycaemia, however the female KO mice required a

significantly lower glucose infusion rate than littermate controls (Figure 3.13B). Such

a finding indicates a greater secretion of glucagon by KO mice. Serum glucagon

measurements performed on samples taken from these mice showed that at time

zero, KO mice had a lower basal release of the hormone but at 90 min after the mice

were rendered hypoglycaemic they released significantly more glucagon than WT

mice (Figure 3.13C).

Figure 3.13 – Hypoglycaemic hyperinsulinaemic clamps. These were performed on 12 week

old mice. They were rendered hypoglycaemic by administration of insulin (A). Glucose levels

were monitored and glucose infusion rate adjusted accordingly to keep the mice in a

hypoglycaemic state (B). (C) Plasma glucagon was measured from blood samples taken at 0

and 15 min. n=4-6. Data are mean ± SEM, two-way repeated measures ANOVA or two-way

ANOVA. *P<0.05, **P<0.01

Page 91: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

91

3.4 Discussion

ZnT8 is a Zn2+ efflux transporter localized on the secretory granule membrane of β-

but also α-cells where it is believed to be involved in the storage of Zn2+ into the

granule from the cytoplasm [213]. In this part of the study we showed that ZnT8 is

not needed in the α-cell for normal glucose homeostasis or insulin sensitivity

although if a greater degree of deletion was achieved the findings could potentially

have been different. However, its absence from a subset of these cells causes an

exaggerated in vivo response in females under hypoglycaemic conditions imposed by

hypoglycaemic clamps.

It should be noted that mice bearing only the Cre transgene displayed normal

glycaemic control as assessed by fasting/refeeding studies and so were α-cell

responses to various regulators of secretion [297]. Similarly mice bearing only RFP

did not display any obvious defects and had normal fertility [293].

3.4.1 Recombination in the α-cell

FACS data from RFP+ dissociated islets stained for glucagon, insulin and ZnT8

revealed that the recombination efficiency of this mouse model is quite low. Thus,

based on the numbers of cells stained both for glucagon and ZnT8, there was only a

~15% reduction in ZnT8 expression in the α-cell population (Figure 3.6D), while there

was no significant deletion in the β-cell population. Thus although PPGCre appeared

to highly specifically delete ZnT8 in the α-cell and not β-cells or in other glucagon

expressing cells such as the gut or brain [280], this was only achieved in a minority of

α-cells. A low recombination rate has been described by others using Cre expressed

by the same PPG promoter. One report made similar observations to our own with

13% recombination [298], while others saw 35% [299] and 43% [300] deletion.

Page 92: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

92

3.4.2 Glucose homeostasis and insulin sensitivity

Glucose tolerance tests assess the ability of the body to deal with glucose loads.

Given that ZnT8 was deleted only in the α-cell any differences in the KO animals can

reasonably be attributed to an altered capacity of the α-cell to deal with glucose.

Previous published work showed that the response of αZnT8 KO mice to glucose

during IPGTTs was similar to controls [272]. However, this was done only at one age

so testing the mice at a range of ages was deemed necessary to make sure no age-

related differences were missed out. Thus the mice were tested at the ages of 8, 12

and 16 weeks but no apparent differences were observed (Figure 3.9).

Fasting blood glucagon was similar in WT and KO which explains why there were no

differences in fasting blood glucose either (Figure 3.10).

Various groups have worked on mice with global deletions of ZnT8 and in terms of

glucose tolerance as assessed by IPGTTs. Two groups reported no change [268, 270]

while three mouse colonies were reported as glucose intolerant two of which only

displayed this effect at a specific age [222, 269]. Deleting ZnT8 in the β-cell caused

glucose intolerance in only one of the two studies [271, 272].

Insulin sensitivity assessed by insulin tolerance tests did not appear to be altered

either in αZnT8 KO mice (Figure 3.11). However, although blood glucagon levels

were similar, the blood glucose concentration did not go below the threshold

required for stimulation of glucagon secretion in either WT or KO mice. Thus what

we were seeing was not α-cells at their most active. This could be why no difference

was observed in this setting. By contrast in conditions of true hypoglycaemia

occurring during hyperinsulinaemic clamps the KOs behaved differently to the WTs.

Higher insulin concentrations were also tried in ITTs in order to increase the degree

of hypoglycaemia further. However this often resulted in a dramatic decrease of

glucose concentration which made necessary the administration of glucose or even

Page 93: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

93

death as the glucose levels dipped rapidly. Thus these mice could not be included in

the results and a lower insulin level was deemed preferable.

3.4.3 Response to hypoglycaemia in vivo

Clamp studies have shown that female mice in a state of hypoglycaemia secrete 2-

fold more glucagon compared to control mice (Figure 3.13). It is interesting to note

that when ZnT8 was deleted in the β-cell, which functions during glucose

concentrations that are inhibitory to the α-cell, during stimulatory conditions the

opposite had been reported. That is insulin secretion in vivo is hindered by the

absence of ZnT8 [271]. Thus, deleting ZnT8 in these two cell types, which have

contrasting functions, causes an effect in the opposite direction.

3.5 Conclusion

Based on the findings of this study ZnT8 is not required in the α-cell for normal

glucose homeostasis. However this transporter is necessary in a subset of the α-cell

population normal responses to hypoglycaemia in vivo.

Page 94: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

94

Chapter 4

ZnT8 is necessary in the α-cell for normal cytosolic and

granular Zn2+ homeostasis

Page 95: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

95

4.1 Introduction

Zinc is one of the most important transition metals in the human body. In its ionic

form Zn2+ is an essential component of many proteins and is vital for a multitude of

cellular processes including cell division, transcription and signalling [301]. The Zn2+

content of the pancreas is one of the highest in the body [233] consistent with the

findings that these ions have a key role in insulin storage and secretion in β-cells

[266].

Diabetes has been known for many years to have an impact on Zn2+ homeostasis in

the body. Thus published work has shown that compared to healthy controls T2D

patients have a substantial reduction in total Zn2+ both in blood and at the cellular

level [236]. Similarly, murine genetic models of T2D exhibit lower pancreatic Zn2+

levels [238, 239].

Insulin in secretory granules is packaged in a hexameric structure co-ordinated by

two Zn2+ ions [267]. Correspondingly, autometallography has shown that Zn2+ in β-

cells is found in close association to secretory granules, with variable staining ranging

from heavy, sparsely or non-stained granules [302]. Although less intense the

staining was also present in α-cell secretory granules, mainly in the periphery, and all

α-cells examined under the electron microscope were stained to an extent [302].

Thus, the presence of Zn2+ in α-cell secretory granules implies that it could serve an

important function in the regulation of glucagon secretion and not just insulin

secretion. The fact that the Zn2+ transporter ZnT8 has been found to be encoded by a

T2D associated gene [203] further reinforces the likelihood of Zn2+ playing an

important role in islet cells. However it has generally been assumed that the role of

Zn2+, and thus of ZnT8 variants, in altering diabetes risk is mediated via alteration in

insulin release.

Page 96: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

96

Many studies either on global or β-cell specific ZnT8 KO mice, have produced

controversial findings as to its role in the β-cell (see Introduction, chapter 1.13.1).

Although some groups have shown no differences in the majority of glycaemic and

other parameters tested, notably glucose tolerance and glucose-stimulated insulin

secretion, others have found that the absence of ZnT8 had a significant impact. This

included impaired glucose tolerance [222, 269], insulin secretion in vivo [222, 271],

reduced intracellular Zn2+ accumulation [222, 268] and morphological changes to

granule infrastructure [222, 270].

The α-cells of the pancreatic islet are electrically excitable, using electrical signals to

couple changes in plasma glucose levels to altered glucagon secretion, responding to

direct and indirect stimuli. Several candidates have been suggested to affect

glucagon secretion in a paracrine manner, insulin and other β-cell products such as

Zn2+ being some of these candidates. The cellular compartment with the highest

amount of Zn2+ in β-cells is the secretory granule, containing up to 70% of total Zn2+

and reaching a concentration of ~10-20 mM [129]. Upon exocytosis, due to the

increase in pH from ~5.5 in granules to ~7.4 in blood, the insulin crystals become

destabilized releasing both Zn2+ ions and creating a large local concentration of Zn2+.

Various studies have therefore raised the possibility for Zn2+ acting as a paracrine

regulator of glucagon secretion. Studies on rodent islets and cells have shown an

inhibitory effect of Zn2+ on glucagon release [121, 130].

Given the findings that ZnT8 appears to be important in the α-cell for the normal

regulation of glucagon secretion (Chapter 3) the impact of the absence of ZnT8 on

glucose control of signalling to both Zn2+ and Ca2+ in the α-cell was investigated.

Page 97: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

97

4.2 Aims

Here we wanted to examine the effects of deleting ZnT8 in the α-cell at the single

cell level particularly to check whether this transporter is important in regulating

cellular Zn2+ homeostasis as observed in the β-cell. This was feasible using mice with

RFP-labelled α-cells for imaging live cells or identification for other in vitro

experiments.

Page 98: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

98

4.3 Results

4.3.1 Impact of α-cell ZnT8 deletion on glucagon secretion in vitro

Mice lacking ZnT8 selectively in the α-cell were generated and maintained as

described in Chapter 3. The deletion of Slc30a8 specifically in the β-cell has

previously been shown to influence in vitro insulin secretion [252]. To determine

whether islets from αZnT8 KO mice respond differently to those from WT animals

glucagon secretion experiments were done on freshly isolated islets. WT and KO

islets were incubated with 1 mM or 17 mM glucose for 1 h and the glucagon released

was quantified using a radioimmunoassay. There was no change in secretion at 17

mM but at stimulatory conditions of 1 mM glucose, KO islets secreted significantly

more glucagon compared to controls (Figure 4.1).

Figure 4.1 - Glucagon secretion in vitro.

Isolated islets were incubated in 1 mM

and 17 mM glucose for 1hr. The

glucagon was the measured using an

immunoassay. The KO islets secreted

significantly more glucagon compared to

control. Data are mean ± SEM, n=5,

**P<0.01

4.3.2 Calcium responses

Differences in secretion capacity at the level of the individual α-cell might be caused

by altered intracellular free Ca2+ changes in response to low glucose concentrations

[303]. Thus, the impact of glucose on α-cell intracellular free Ca2+ levels was

investigated in whole islets by intracellular Ca2+ imaging with Fluo-2, an

intracellularly-trappable fluorescent Ca2+ indicator (Figure 4.2A). RFP+ WT and KO

islets were used in this instance for identification of α-cells and these were imaged

during exposure initially to 1mM then to 11 mM glucose (Figure 4.2B,C). The peak

1 1 7

0 .0 0

0 .0 5

0 .1 0

0 .1 5

0 .2 0

0 .2 5W T

K O

**

G lu c o s e (m M )

Glu

ca

go

n r

ele

as

e %

to

tal

Page 99: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

99

Ca2+ levels apparent in the traces obtained (AUC; Figure 4.2D) and the area under

the curve (Figure 4.2E) showed that Ca2+ responses were not altered in the KO islets

compared to controls. Similarly, Fourier Transform analysis of the traces showed no

differences in the average frequency of oscillations (Figure 4.2F). Thus, exaggerated

Ca2+ responses to glucopenia cannot explain the increased glucagon secretion at 1

mM. Additionally we performed the statistical D’Agostino Pearson omnibus

normality test (Graph Pad Prism) which both WT and KO islets passed, showing that

the distribution of frequencies was Gaussian and there was no evidence for the

existence of two populations of cells.

Figure 4.2 – Low glucose-induced calcium responses. RFP+ WT and KO islets were incubated

with the calcium indicator Fluo2 (A) for 40 min before imaging at 1mM and 11mM glucose.

Scale bar 50 µm. Representative traces from WT (B) and KO (C) α-cells. (D) Area under the

curve (AUC). (E) Spike amplitude as a % of baseline. (F) Fast Fourier transform of calcium

traces. Data are mean ± SEM, n = 16-18 islets from 5 mice/genotype.

0.0

0

0.0

1

0.0

2

0.0

3

0.0

4

0.0

5

0.0

6

0.0

7

0.0

8

0

2 0

4 0

6 0

8 0

K O

W T

D o m in a n t fre q u e n c y (H z )

Pe

rce

nta

ge

ce

lls

(%

)

T im e (m in )

F/F

min

0 1 0 2 0

1 .2

1 .6

2 .0

2 .4

G 1 G 11

T im e (m in )

F/F

min

0 1 0 2 0

1 .2

1 .6

2 .0

2 .4

G 1 G 11

W T K O

0

2 0 0

4 0 0

6 0 0

8 0 0

1 0 0 0

AU

C (

AU

)

W T K O

1 0 0

1 2 0

1 4 0

1 6 0

1 8 0

Sp

ike

am

pli

tud

e R

FP

+v

e c

ell

s

(% o

f b

as

eli

ne

)

RFP Fluo-2 A C B

D E F

Page 100: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

100

4.3.3 α-cell area

One possible driver of increased glucagon secretion would be an increase in

pancreatic α-cell mass, and/or a change in the proportion of α- versus β-cells. To

assess this possibility, pancreatic slices from the above mice were stained for

glucagon and insulin (Figure 4.3A) and imaged to determine the α- and β-cell area. In

KO mice (Figure 4.3B) there was a tendency for increased α-cell area and islets from

α-ZnT8 KO mice displayed a significant increase in α-to-β cell area ratio.

Figure 4.3 – Pancreatic histology. (A) Slices from fixed WT and KO pancreata were stained

for insulin (green) and glucagon (red). (B) The relative area is shown as α:β cell ratio. Scale

bar 50 µm. Data are mean ± SEM, n=5 mice, 8-10 slices per mouse, *P<0.05

4.3.4 Intracellular Zn2+ imaging

Given its role in the β-cell [222], where ZnT8 is involved in the transfer of Zn2+ from

the cytoplasm into the granule, we anticipated that a similar role in the α-cell would

mean that its absence could have an effect on cytoplasmic Zn2+ levels. In our lab Zn2+

imaging in islets is routinely done using a recombinant Zn2+ probe (eCALWY-4)

packaged in adenoviral particles [288]. Initially this probe was used in αTC1.9 cells, a

mouse α-cell line, to determine the free cytosolic Zn2+ concentration in the α-cell.

The cells were first imaged without additions to get a baseline and then after adding

TPEN, a membrane permeable chelator, followed by excess Zn2+ with the ionophore

pyrithione (Figure 4.4A). This calibration allows the determination of cytoplasmic

W T K O

0 .0 0

0 .0 5

0 .1 0

0 .1 5

0 .2 0

0 .2 5 *

:

-ce

ll r

ati

o

A B

Glucagon Insulin

WT KO

Page 101: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

101

0 2 0 0 4 0 0 6 0 0 8 0 0

3 .5

4 .0

4 .5

5 .0

5 .5

T P E NZ nP yr

T im e (s e c )

Cit

rin

e/C

eru

lea

n

6 0 0

7 0 0

8 0 0

9 0 0

1 0 0 0

[Zn

2+] c

yt

(pM

)

A B

[Zn2+] for each individual cell using the changes in FRET ratio which occur after the

aforementioned additions. Results from imaging with this probe (Figure 4.4B) on a

mouse α-cell line showed that cytoplasmic [Zn2+] in mouse α-cells ranged from 680-

920 pM averaging ~810 pM.

Figure 4.4 – Zn2+ imaging with eCALWY-4. αTC1.9 cells were infected with eCALWY-4

adenovirus bearing the cytosolic form of the probe. Cells were imaged the next day first at

baseline followed by the additions indicated on the representative trace (A). (B) Cytosolic

[Zn2+] values from individual α-cells.

Adenoviruses are very efficient at infecting primary β-cells but appear to be very

inefficient in the case of α-cells. This has been seen in our laboratory [304] and by

others [305] but was also observed by us after staining of fixed islets for glucagon

following infection of whole islets with eCALWY-4. Normally a multiplicity of

infection (MOI) of 20-25 [306] is used in experiments of this sort. In an effort to

assess whether a higher number of viral particles could be effective in driving

infection of α-cells, MOIs of 500, 2500 and 5000 were used for infection of islets

prior to fixation. Imaging of islets infected in this way showed that although there

was good infection of the islet as a whole, this only represented the β-cell population

since the glucagon-positive cells were still largely negative for the GFP-containing

virus (Figure 4.5). Increasing the MOI by 20 or even 200 fold did not appear to

detectably improve the infection of α-cells (Figure 4.5A-C).

Page 102: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

102

Figure 4.5 – Islet expression of adenoviral Zn2+ probe. Islets infected with eCALWY-4

adenovirus were fixed and stained for glucagon. MOIs of 500 (A), 2500 (B) and 5000 (C) were

used to check if an increased viral load affected the efficiency of infection of α-cells.

In an effort to overcome this issue of very low infectivity by adenoviruses we sub-

cloned the virus into an adeno-associated virus (AAV) potentially allowing uptake by

a distinct set of cell surface receptors. In this case the viral genome had a flex switch,

so the coding sequence was placed in the opposite orientation to that required for

transcription. In the presence of Cre, which would recognize the lox sites, this

orientation would be flipped allowing expression of eCALWY4 (Figure 4.6A). This

would have been ideal for expression of the probe in the primary α-cells in our

mouse model which express Cre.

Positive clones were identified (Figure 4.6B) and AD293 cells were co-transfected

with a Cre-expressing plasmid showing robust expression of eCALWY-4, confirming

eCALWY-4 glucagon merge A

C

B

Page 103: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

103

eCALWY-4 2.2kb

2Kb

A

C B

that the probe was functional in the AAV construct (Figure 4.6C). The virus was then

generated and tested but we could not detect expression of the probe in islets

positive for Cre in the α-cells nor in β-cells, even when using other AAVs containing

other recombinant proteins which were shown to work when used in the brain. It

thus seems that the AAV serotypes 1 and 5 harboured by the virus bearing the probe

are not effective at infecting α-cells or β-cells thus making this virus unusable in

primary islet cells.

Figure 4.6 – AAV-eCALWY-4. (A) DNA construct and recombination events for Cre-

dependent eCALWY-4 expression. (B) Double RE digest (SpeI, EcoRI) of positive colonies

containing the construct. Lanes 1, 3, 5-undigested DNA. Lanes 2, 4, 6-double digested DNA.

(C) Image of AD293 cells expressing eCALWY-4 following transfection with a Cre-pCDNA3.1

and eCALWY-4-pAAV plasmids

Since the adenoviral or AAV recombinant probe could not be utilised an alternative

was found, a loadable Zn2+ indicator FluoZin-3, which is expected primarily (though

perhaps not exclusively) to be localized to the cytoplasm [307]. RFP+ WT and KO

islets were incubated with the dye for an hour before imaging (Figure 4.7A) initially

perfusing with no additions to allow a recording of baseline fluorescence. At about

10 min the Zn2+ chelator TPEN was added, binding essentially all intracellular labile

Zn2+ and causing a reduction in fluorescence of the dye (Figure 4.7B). The

Page 104: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

104

fluorescence value at the baseline and the minimum value following the addition of

TPEN were used to give a ratio which gave an indication of the free Zn2+

concentration in the cytoplasm. Note that the maximum fluorescence in the

presence of high intracellular Zn2+ concentrations after the addition of ZnCl2 to the

cells, in the presence of the ionophore pyrithione could not be used to provide a

precise calibration of fluorescence in terms of free Zn2+ it caused a saturation of the

signal (not shown). Nevertheless, based on the values determined as above,

intracellular Zn2+ levels appeared to be lower in KO compared to WT α-cells (Figure

4.7C). A similar difference was observed previously in whole islets from β-cell specific

KO animals, in this case using the recombinant probe eCALWY4 [222]. No such

difference was evident in non-RFP labelled cells which were largely made up of the

β-cell population of the islet, and in which ZnT8 was still expressed (Figure 4.7D).

Figure 4.7 – Intracellular Zn2+ imaging. RFP+ WT and KO islets incubated with the Zn2+

indicator FluoZin-3 for 1 h (A) were imaged at baseline (F1) and following perfusion with the

Zn2+ chelator TPEN (Fmin). (B) Representative trace from an α-cell. (C) Zn2+ levels of RFP+ cells,

quantified from the relative fluorescence (F1/Fmin). (D) Zn2+ levels of RFP- cells. Data are mean

± SEM, n = 16-20 islets from four mice, Student’s t-test, ***P<0.0001

W T K O

1 .0

1 .5

2 .0

2 .5

F1

/Fm

i (A

U)

0 1 0 2 0 3 0

3 0 0 0

4 0 0 0

5 0 0 0

6 0 0 0

7 0 0 0

8 0 0 0

T P E NF 1

F m in

T im e (m in )

Flu

ore

se

nc

e (

AU

)

W T K O

1 .0

1 .5

2 .0

F1

/Fm

in (

AU

)

****

RFP FluoZin-3

A

C

B

D

Page 105: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

105

4.3.5 Impact of α-cell ZnT8 deletion on secretory granule Zn2+ content

Considering the proposed location of ZnT8 on secretory granules and its implied role

in maintaining granular Zn2+ concentration the next step was to search for any

changes in granular Zn2+ content. For this the RFP+ WT and KO islets were incubated

with the loadable Zn2+ dye Zinpyr4 [283]. Experiments were first done to check that

it colocalized with the granules. INS-1 (832/13) cells were loaded simultaneously

with the dye and a fluorescent Lysotracker [308] with an orthogonal emission

spectrum. The latter, although usually specific for lysosomes in non-secretory cells,

also enters insulin granules due to their mildly acidic environment and positive

charge with respect to the cytosol [309]. As shown in Figure 4.8A the staining not

only appeared granular but there was also good colocalization between Zinpyr4 and

the lysotracker, verifying the localization of the Zn2+ probe to the granule interior.

However, this dye photobleached very quickly even at the low laser powers used so

a decrease in signal was evident in real time when attempting to make a long

acquisition. It was thus not possible to acquire a similar trace as the one obtained

with the cytoplasmic dye FluoZin-3 (Figure 4.7B). Therefore, imaging of Zinpyr4 was

restricted to the acquisition of static images (Figure 4.8B) whilst all the conditions

such as exposure time, laser power and incubation time were kept the same to make

sure any variance in fluorescence was not a result of these variables being different.

The fluorescence intensity was measured using the corrected total cell fluorescence

equation (CTCF), a parameter which eliminates any discrepancy in actual

fluorescence due to cell size. Compared to RFP+ cells from WT islets, KO cells

exhibited an approximated 50% reduction in apparent free granular Zn2+ (Figure

4.8C).

4.3.6 Gene expression analysis

To confirm the deletion of ZnT8 at the mRNA level and to establish whether there

were any changes in the expression of other genes in the ZnT family, WT and KO

islets were isolated, dissociated and then sorted through a flow cytometer based on

RFP fluorescence. RNA was extracted from the RFP+ α-cells collected during the sort

Page 106: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

106

and expression levels of ZnTs and MTs were assessed by qPCR. As previously shown

by FACS ZnT8 was not completely absent from the RFP+ population, with expression

in KO cells reaching about 40% that of WT cells (Figure 4.9A). Purified α-cells

expressed low levels of ZnT1, 4-7 and 9 while ZnT2 and 3 were not detectable (Figure

4.9A). Deleting ZnT8 caused no major changes in the expression levels of these

genes. WT cells also expressed the cytoplasmic Zn2+ binding proteins MT1 and MT2

but not MT3 with levels being comparable in WT and KO mice (Figure 4.9B). We also

looked into the expression levels of KATP channel subunits and Gck, in order to

determine whether there were changes in the energy sensing components of the

cell. Likewise we examined the expression of Arx and Pcsk2, whose gene products

are involved in α-cell identity and glucagon processing respectively (Figure 4.9C).

There were no changes in the expression levels of these genes nor in the expression

of glucagon in ZnT8 null α-cells (Figure 4.9D).

Figure 4.8 – Granular free Zn2+ imaging. (A) INS1 cells were incubated with ZinPyr-4 and

Lysotracker to determine the localization of ZinPyr-4 in secretory granules. (B) RFP+ WT and

KO islets were incubated with the Zn2+ dye Zinpyr-4 for 30 min before imaging Quantification

of fluorescence is shown as CTCF (Corrected Total Cell Fluorescence) (C) calculated using the

area of the cell chosen, integrated density and background fluorescence. Scale bar 10 μm.

Data are mean ± SEM, n=15-20 islets from three mice, Student’s t-test, **<0.002

W T K O

0

4 .01 0 5

8 .01 0 5

1 .21 0 6

CT

CF

**

RFP Zinpyr-4

KO

WT

Lysotracker Zinpyr-4 A

C B

D

Page 107: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

107

Figure 4.9 – Gene expression analysis. RFP+ WT and KO islets were dissociated and then

sorted. The FACS sorted RFP+ cells were used for RNA extraction and qPCR. (A) Expression

levels of the ZnT family of transporters, (B) metallothioneins, (C) energy sensing and α-cell

specific genes, (D) glucagon in WT and KO cells. n= 5, ***P<0.001

4.3.7 Glucagon content

Increased glucagon content in KO α-cells might explain the enhanced glucagon

secretion observed at stimulatory glucose conditions. The possibility was examined

using two ways. One was to isolate pancreata from WT and KO mice and extract

proteins for total glucagon measurement per unit tissue with an HTFR assay. At the

pancreatic level no differences were observed in the glucagon content between WT

and KO mice (Figure 4.10C). The second method was to isolate RFP+ islets and stain

for glucagon following fixation. Then, the glucagon content of the RFP+ cells was

imaged by taking z-stacks which were then used to measure the total intracellular

glucagon volume and thus the content (Figure 4.10A). Again there were no

differences in total content comparing WT to KO cells (Figure 4.10B).

M T 1 M T 2 M T 3

0 .0

0 .1

0 .2

0 .3

0 .4

0 .5

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

ac

tin

)

W T K O

0

5 0

1 0 0

1 5 0

2 0 0

2 5 0

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

ac

tin

)

n s

A rx A b c c 8 K c n j1 1 G c k P c s k 2

0

1

2

3

4

W T

K O

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

ac

tin

)

Z n T 1 Z n T 2 Z n T 3 Z n T 4 Z n T 5 Z n T 6 Z n T 7 Z n T 8 Z n T 9

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

W T

K O

* * * *

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

ac

tin

)

A

C

B

D

Page 108: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

108

Figure 4.10 – Glucagon content measurements. (A) WT and (B) αZnT8 KO islets were

stained for glucagon and imaged on a confocal microscope at 63x magnification. (C) RFP+

cells were selected for z-stacks and their glucagon volume was measured using ImageJ.

n=45-50 cells from 3 animals. (D) Pancreata from WT and KO mice were isolated and protein

was extracted using an acidic ethanol method and glucagon measured using an HTRF assay.

n= 3

W T K O

0 .0

0 .5

1 .0

1 .5

2 .0

Glu

ca

go

n c

on

ten

t

(n

g/m

g p

an

cre

as

)

W T K O

0

2 0

4 0

6 0

8 0

Glu

ca

go

n c

on

ten

t (

m3

)

A B

C D

Page 109: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

109

4.4 Discussion

In this chapter we investigated the effect of ZnT8’s absence on α-cell in vitro and at

the single cell level. Work on isolated islets confirmed the finding of enhanced

glucagon secretion during hypoglycaemia (Figure 4.1). Zn2+ imaging showed a

reduction in free Zn2+ levels both intracellularly and in the secretory granules.

Additionally, we showed that the increased glucagon secretion observed during

hypoglycaemia both in vivo and in vitro cannot be explained by increased glucagon

content or by heightened Ca2+ responses to low (stimulatory) glucose concentrations

in the KO α-cell.

4.4.1 Glucagon secretion

Glucagon release from isolated islets at 1mM glucose was ~80% higher from KO

islets compared to WT controls (Figure 4.1). This is in agreement with the in vivo

findings of enhanced glucagon secretion in hypoglycaemic conditions (Figure 3.13).

Imaging stained pancreatic slices showed an increase in the α:β-cell ratio by ~15%

which could partly contribute to the increased glucagon secretion at low glucose.

However changes in some other cellular mechanisms must be taking place in the KO

mice that could explain this large increase in glucagon release.

Zn2+ homeostasis in the α-cell

ZnT8 was found to be important in Zn2+ homeostasis in the β-cell [222] and thus we

wished to investigate whether these ions played a similar role in the α-cell. First, we

examined cytosolic Zn2+ levels using a protein-based Zn2+ indicator. The recombinant

Zn2+ probe would have been preferable since it would be strictly targeted to the

cytosol and also it would allow us to use the ratio of fluorescence from the two

fluorophores it contained. Calibrations based on its dissociation constant for Zn2+

would then have allowed us to obtain values in molar terms for the free

concentration of Zn2+. Unfortunately, however, the traces obtained with the

Page 110: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

110

alternative FluoZin-3 probe could not be calibrated accurately due to the dynamic

range of the indicator which meant that the maximum fluorescence was not

detected properly (i.e. the signal was saturated). Consequently these could not be

used in a calculation for Zn2+ concentration.

Despite this limitation we were able to use the FluoZin-3 fluorescence values at

baseline, and after the cell was depleted of Zn2+, to give a ratio which suggested that

Zn2+ levels in the cytosol were reduced in KO vs WT cells. Such a finding although

consistent with observations from the β-cell [310], are nonetheless somewhat

surprising. The lack of an efflux transporter is not expected to have resulted in a

reduction in cytosolic Zn2+. In fact, the change should have been in the opposite

direction. This indicates that ZnT8 may in normal conditions be acting as an influx

transporter at sites other than the granules such as the plasma membrane. Similar

findings have been made for another member of the ZnT family, ZnT5 [311].

Similarly, using ZinPyr-4 we imaged granular Zn2+ content. In this case the findings

were consistent with the proposed function of ZnT8 in transporting Zn2+ across the

granule membrane and into the granule lumen from the cytosol. As expected the

absence of ZnT8 led to a reduction in granular content (Figure 4.8C) reflecting a loss

of transporting activity in the KO cells.

4.4.2 Absence of effects of ZnT8 inactivation in α-cells on glucose-regulated Ca2+

fluxes

Interestingly, the absence of ZnT8, and consequent changes in cytosolic and granule

Zn2+, exerted no apparent effects on Ca2+ dynamics in α-cells, despite reports of

multiple interactions between these two ions on Ca2+, K+ and other channels [312].

Further, more detailed investigations possibly involving electrophysiology may be

needed to explore this surprising result in the future.

Page 111: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

111

4.5 Conclusion

ZnT8 appears to be important in Zn2+ homeostasis in the α-cell influencing both

cytosolic and granular compartments leading to a reduction in concentration which

cannot be explained by changes in any of the other zinc transporters examined. This

direction of change in cytosolic concentration points towards a role for ZnT8 in Zn2+

influx and not just efflux. A mechanism other than enhanced calcium responses must

be responsible in causing the enhanced glucagon secretion during hypoglycaemia

possibly involving Zn2+.

Page 112: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

112

Chapter 5

Overexpression of human ZnT8 in the murine pancreatic

α-cell leads to reduced glucagon release and impaired

responses to hypoglycaemia

Page 113: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

113

5.1 Introduction

Functional studies in model systems are vital if we are fully to understand the

physiological role(s) of genes identified through genome-wide association and other

studies as affecting the risk of T2D [212]. Most studies up to now have used gene

deletion globally or in disease-relevant tissues to explore this question (see

Introduction, and Chapters 3,4) [210, 212]. Although investigating the absence of a

gene can be very informative its overexpression can likewise give similarly important

findings into the role of a protein. In the case of SLC30A8/ZnT8 this is of particular

importance since the impact of risk variants on activity and diabetes risk is still open

to discussion: the common risk variant (R325) is thought to lower activity and

increase risk [222], whereas rare truncated variants appear to be protective [221].

Inducible expression systems are commonly used in rodent systems to achieve both

temporal and spatial (i.e. tissue-specific) expression of a given gene. Components of

the Tet Switches [313] originate from the tetracycline (Tet) resistance operon in

E.coli belonging to one of the most evolved gene regulation systems. The Tet-Off and

Tet-On systems are used in the majority of the studies involving inducible

expression. The Tet-Off system was the initial system developed in 1992 and in the

presence of the antibiotic tetracycline the expression from a tet-inducible promoter

is decreased.

In order to utilize tetracycline as a regulator of transcription, a tetracycline-

controlled transactivator (tTA) was made by fusion of the tetracycline repressor with

a transcriptional activation domain from HSV. Thus, in the absence of tetracycline

the fusion protein can bind tet operator sequences and promote transcription while

in the presence of the antibiotic, its binding to the protein makes it unable to bind

DNA leading to a decrease in gene expression. The Tet-On system was later

developed by mutation of the repressor portion of the tTA so that it would rely on

Page 114: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

114

tetracycline for induction of gene expression and not repression. This is known as a

reverse tetracycline controlled transactivator (rtTA) [314].

The system was first used in the β-cell [315] and about 10 years later the system was

utilised in the α-cell [316]. Recently, our laboratory has used this approach to

determine the effects of ZnT8 over-expression in the pancreatic β-cell in mice by

driving rtTA expression with the rat insulin promoter [317]. In the present study, the

rtTA sequence was placed under the control of the PPG promoter to generate

GlurtTA mice and to drive expression of this gene selectively in the α-cell in the adult

mouse.

5.2 Aims

Here, we wanted to investigate the effect the overexpression of ZnT8 would have on

glucagon secretion, GlurtTA mice were therefore crossed to mice bearing a human

ZnT8 transgene whose expression was driven by a tet operator sequence.

Page 115: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

115

5.3 Results

5.3.1 Generation and genotyping of αZnT8Tg mice

For this project, GlurtTA mice were obtained (Dr Ahmed Mansouri, Max-Planck

Institute, Goettingen, Germany) and crossed to animals containing a human ZnT8-

overexpressing transgene (ZnT8Tg) as described [317]. Heterozygote GlurtTA mice

were crossed to homozygote ZnT8Tg animals. Two founder were used,

corresponding to lines #31 and #23 in Mitchell et al [317]. The litters comprised

GlurtTA offspring, which expressed the human ZnT8 transgene after induction with

doxycycline (as described above), while the littermates bearing the ZnT8Tg allele

alone were used as controls.

GlurtTA mice possess a 1.1kb region of the glucagon promoter upstream of the rtTA

[316]. In the α-cell the transactivator is produced but is not in its active form.

However doxycycline, a tetracycline derivative, binds and activates it so that it can

then in turn interact with the Tet operator sequence upstream of the construct

containing the cDNA of human ZnT8 and transactivate its expression. Transcription is

only possible as long as doxycycline is present. Consequently, the drug needs to be

administered continuously [316]. Mice were given doxycycline in the drinking water

(2g/L) from the age of 6 weeks.

The genotyping for GlurtTA was performed using standard PCR on DNA extracted

from ear biopsies (Figure 5.1). For the ZnT8 transgene, qPCR was used to detect the

presence in the genome of the luciferase gene, also included in the transgene.

Figure 5.1 – Genotyping by PCR. The

GlurtTA transgene was amplified by

standard PCR to identify control and

transgenic animals. Lanes 1, 2, 4, 6, 7, 9,

11 contain samples from transgenic

animals.

200 bp

Control

GlurtTA

1 2 3 4 5 6 7 8 9 10 11

Page 116: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

116

5.3.2 Expression of human ZnT8 in the islet

Initially to check that the administration of doxycycline was effective in inducing the

overexpression of ZnT8, total RNA was extracted from isolated islets incubated

overnight with doxycycline (5µg/ml). qPCR analysis revealed expression of the

human form of ZnT8 in the mice containing both the overexpressing transgene and

the GlurtTA but not in the control mice carrying only the ZnT8 transgene (Figure 5.2).

We tested both founders that contained the transgene which showed that only

founder 2 (#31 from Mitchell et al [317]) expressed detectable levels of human ZnT8.

Consequently, this founder was used in all experiments.

Figure 5.2 – Human ZnT8

expression in isolated islets.

Total RNA was extracted from

islets from control and

transgenic mice. The levels of

hZnT8 were assessed by RT-

qPCR.

In order to obtain a more accurate assessment of the degree of overexpression in a

pure α-cell population and as opposed to islets, which are largely composed of β-

cells [318], isolated islets were dissociated, fixed and stained with glucagon for

fluorescent –activated cell sorting. RNA was then extracted from the sorted cells to

check for expression of the human ZnT8 in the α-cells obtained from the transgenic

and control animals (Figure 5.3).

c o n tro l fo u n d e r 1 fo u n d e r 2

0 .0 0 0

0 .0 0 1

0 .0 0 2

0 .0 0 3

mR

NA

ex

pre

ss

ion

(re

lati

ve

to

-ac

tin

)

Page 117: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

117

Figure 5.3 – Hum an ZnT8 expression in

purified α-cells. RNA was extracted from

α-cells obtained from control or αZnT8Tg

mice. The levels of hZnT8 mRNA were

assessed by RT-qPCR. n=3

5.3.3 Intraperitoneal glucose tolerance was unaltered in α-ZnT8Tg mice

To check glucose homeostasis in these mice, IPGTTs were performed at the age of 8

weeks (Figure 5.4A) and 12 weeks (Figure 5.4B). Similarly to our findings from αZnT8

KO animals (Chapter 3) there was no difference in blood glucose between control

and transgenic animals at any time point during the test.

Figure 5.4 – Glucose tolerance in αZnT8Tg mice. Intraperitoneal glucose tolerance tests

were performed on control and transgenic mice at 8 (A) and 12 weeks of age (B). n=8-10

mice/genotype

0 3 0 6 0 9 0 1 2 0

0

5

1 0

1 5

Glu

co

se

(m

M)

C o n tro l

T ra n s g e n ic

0 3 0 6 0 9 0 1 2 0

0

5

1 0

1 5

T im e (m in )

Glu

co

se

(m

M)

C o n tro l

T ra n s g e n ic

A B

C o n tro l T ra n s g e n ic

0 .0

0 .2

0 .4

0 .6

0 .8

hZ

nT

8 m

RN

A e

xp

res

sio

n

(re

lati

ve

to

-ac

tin

)

Page 118: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

118

5.3.4 Intraperitoneal insulin tolerance tests (ITTs)

ITTs were performed on transgenic and control male (Figure 5.5A) and female mice

(Figure 5.5B) at the age of 14 weeks in order to check their insulin sensitivity. The

responses appeared to be the same in the ZnT8 overexpressing and control mice.

5.3.5 Hypoglycaemic clamps

As discussed in Chapter 3, the above experiments could only be used safely to

provide a limited decrease in blood glucose levels (to ~4mM). In order to produce a

more marked decrease in blood glucose levels, where glucagon release is likely to be

more strongly stimulated hypoglycaemic clamps were performed on male mice in

Prof Christophe Magnan’s laboratory (University Paris Diderot-CNRS UMR8251). To

investigate the in vivo response to the imposed low glucose levels the mice were

rendered hypoglycaemic by continuous intravenous administration of insulin (Figure

5.6A) (see Methods, Chapter 2). ZnT8 overexpressing mice required a significantly

higher glucose infusion rate to maintain hypoglycaemia compared to control mice

(Figure 5.6B), suggesting an impairment in glucagon release. Glucagon levels were

consequently measured prior to the onset of hypoglycaemia and at 120 min after the

induction of hypoglycaemia (Figure 5.6C). This revealed lower plasma glucagon levels

at the latter time point, consistent with the higher glucose infusion rates required

during the clamps.

Page 119: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

119

Figure 5.5 – Insulin sensitivity. Intraperitoneal insulin tolerance tests performed on 12

week old male (A) and female (B) mice fasted for 5 hr and injected with 0.5U/kg insulin. n=9-

10 mice per genotype

5.3.6 Glucagon secretion is impaired in islets isolated from αZnT8Tg mice

We sought next to determine whether the impaired glucagon secretion observed in

vivo in αZnT8 mice chiefly reflects a cell autonomous effect in the α-cell, as opposed

to a more complex set of changes in the autocrine response to hypoglycaemia,

potentially involving multiple tissues (brain, adrenals etc) [319]. Furthermore, we

have previously shown that deleting ZnT8 from the α-cell enhances glucagon

secretion from isolated islets (Chapter 3) and were therefore interested in the

hypothesis that overexpression may have an effect in the opposite direction i.e. to

impair glucagon secretion in response to low glucose. Correspondingly, we observed

that when isolated islets were incubated in 1 mM glucose (a concentration in the

range which strongly stimulates glucagon secretion from α-cells) [88], those

overexpressing ZnT8 secreted significantly less glucagon that control islets (Figure

5.7). At 17 mM glucose the responses did not differ between islets overexpressing

ZnT8 selectively in the α-cell and control islets.

0 1 5 3 0 4 5 6 0

0

2

4

6

8

1 0

T im e (m in )

Glu

co

se

(m

M)

C o n tro l

T ra n s g e n ic

0 1 5 3 0 4 5 6 0

0

5

1 0

1 5

T im e (m in )

Glu

co

se

(m

M)

C o n tro l

T ra n s g e n ic

A B

Page 120: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

120

Figure 5.6 – Elevated glucose infusion rates in αZnT8 KO mice during hypoglycaemic

hyperinsulinaemic clamps. Experiments were performed on 12 week old male mice. Animals

were rendered hypoglycaemic by infusion of insulin (A). Glucose levels were monitored and

glucose infusion rate adjusted accordingly to keep the mice in a hypoglycaemic state (B). (C)

Blood glucagon levels were measured at 0 and 120 min. n=4. Data are mean ± SEM, *P<0.05,

**P<0.01, ***P<0.005.

Figure 5.7 - Glucagon secretion from αZnT8Tg

islets is impaired in vitro. Isolated islets were

incubated at 1 mM or 17 mM glucose for 1 h.

Released glucagon was measured using an

HTRF assay. Transgenic islets secreted

significantly less glucagon compared to control

when incubated at 1 mM glucose. Data are

mean ± SEM, n=4 separate experiments,

*P<0.05

1 1 7

0 .0 0

0 .0 5

0 .1 0

0 .1 5

0 .2 0

0 .2 5

C o n tro l

T ra n s g e n ic

G lu c o s e (m M )

Glu

ca

go

n r

ele

as

e %

to

tal *

0 2 0 4 0 6 0 8 0 1 0 0 1 2 0 1 4 0

0

5

1 0

1 5

T im e (m in )

Gly

ca

em

ia (

mM

)

C o n tro l

T ra n s g e n ic

0 2 0 4 0 6 0 8 0 1 0 0 1 2 0 1 4 0

0

2 0

4 0

6 0

T im e (m in )

GIR

(m

g/m

in/k

g)

C o n tro l

T ra n s g e n ic

***

***

0 1200

20

40

60

80Control

Transgenic **

Time (min)

Glu

ca

go

n (p

M)

A B

C

Page 121: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

121

Efforts were next made to detect human ZnT8 by immunostaining using FACS

analysis and confocal microscopy. The antibody used (Mellitech) was described as

specific to human ZnT8. However, analysis of samples by both methods showed that

this might not have been the case. Thus staining of islets did not allow the detection

of human ZnT8 positive populations since there was high background (results not

shown). Similarly when using the antibody to try to identify human ZnT8-positive α-

cells by FACS the entire population of islet cells was found to be positive in both

ZnT8 overexpressing and control mice (results not shown). This was the case each

time regardless of titrating the antibody and using different dilutions to determine if

this was just a result of saturation. Thus, the antibody very likely recognises a mouse

ZnT8 epitope as well as a human one. The alternative ZnT8 antibody available in the

lab was not species specific so could not be of use either.

An anti-c-myc antibody was also tested using cytochemical analysis of isolated islets

to detect the epitope tag also included at the 3’ end of the ZnT8 transgene. However

the staining was very weak with no real positive cells and just a faint background

despite trying several concentrations of the antibody (results now shown).

Page 122: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

122

5.4 Discussion

In this Chapter our aim was to overexpress ZnT8 in the pancreatic α-cell in order to

investigate the effect this would have on glucagon secretion. The results, which are

reminiscent of our findings in αZnT8 KO animals (Chapters 2,3), reveal that

overexpression of ZnT8 in the α-cell has no effect on glucose homeostasis or insulin

sensitivity (assessed by intraperitoneal tolerance tests), but does affect

hypoglycaemia responses as studied in clamps, and glucagon secretion in response

to low glucose both in vivo and in vitro. The findings from the transgenic mouse with

respect to glucagon release are thus reciprocal to those obtained from αZnT8 KO

mice and further reinforce the view that ZnT8 has an important role in the α-cell in

regulating glucagon secretion, at least in the hypoglycaemic state.

The efforts made here to quantify the degree of human ZnT8 expression in the α-cell

at the protein level were, however, unsuccessful. This is, of course, challenging given

that α-cells comprise only a small fraction of the rodent islet [318], requiring single

cell analysis either after FACS or the use of imaging approaches. Had there been

more time, other anti-human ZnT8 antibodies would have been tested in order to

determine the expression of human ZnT8 at the protein level and not just the at the

mRNA level. The fact that expression at the mRNA levels is relatively high compared

to the endogenous control gene lead to the expectation of high enough protein

levels that would be detected by immunohistochemistry. FACS would be even more

sensitive at detecting positive cell populations. It is thus perhaps just an issue of

finding the appropriate antibody.

The results here complement those of a similar study from the laboratory [317]. In

this earlier study, human ZnT8 was inducibly overexpressed in the adult β-cell by

using an insulin promoter-dependent Tet-On system. Glucose-induced insulin

secretion from the βZnT8Tg mice was impaired and Zn2+ release during stimulated

Page 123: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

123

exocytosis was elevated. The same would be expected from the α-cell considering

how deletion of ZnT8 led to a reduction of granular Zn2+ levels in both the α-cell and

β-cell. Further investigations will be required to test this possibility.

In this earlier study, however, the effect of deletion and overexpression of ZnT8 on

insulin secretion were not reciprocal as we observed in the α-cells. Although various

groups [222, 252, 268] have shown an impairment in insulin secretion following ZnT8

ablation, no change was detected in the levels of insulin released. By contrast, Zn2+

release respectively increased and decreased in βZnT8Tg and βZnT8KO mice [317].

5.5 Conclusion

Overexpression of ZnT8 in α-cells, in common with its deletion, affects glucagon

release at hypoglycaemic conditions in this case causing a reduction in secretion.

ZnT8 in the α-cell is thus necessary for normal glucagon responses in the context of

hypoglycaemia. Both rare [221] and common variants [203] in the human SLC30A8

gene may thus lead to changes in both insulin and in glucagon release to affect

glucose homeostasis and T2D risk.

Page 124: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

124

Chapter 6

Final Discussion

Page 125: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

125

6.1 Investigating the role of ZnT8 in the pancreatic α-cell through mouse

molecular genetics

The aim of this thesis was to investigate the role of the Zn2+ transporter ZnT8 in the

α-cell and the regulation of glucagon secretion. The approach here was to generate

mice deleted for or overexpressing ZnT8 specifically in the α-cell. Although ZnT8 has

been widely studied in the β-cell [271, 272] this is the first time that it has been

extensively investigated in the α-cell. The β-cell findings from the various groups

researching ZnT8 have been controversial but strongly imply a role for ZnT8 in β-cell

Zn2+ homeostasis, insulin packaging and insulin secretion. It was thus important to

understand whether ZnT8 is important for α-cell function. A SNP in the SLC30A8

locus encoding ZnT8, resulting in a transporter with reduced transporting ability, was

found to be associated with increased T2D risk [203]. More recently, rare loss-of-

function ZnT8 variants [221] were found to be protective against T2D in contrast to

the previous findings [203]. Understanding the function of ZnT8 not only in the β-cell

but also in the α-cell where it is highly expressed may thus provide some insight and

help determine how and why specific ZnT8 variants affect T2D risk.

T2D is a polygenic disease with many genes possibly contributing to disease

susceptibility and pathogenesis. Gaining knowledge about the role of T2D associated

genes in pancreatic islets could lead to the identification of potential targets for the

development of anti-diabetic drugs.

The ability to generate cell type-specific and whole body knockout mice has made

these an instrumental component in studies investigating the role of a protein in the

whole body or in a certain cell type. Results from rodent studies can be used to

deduce the likely effects a disease-associated gene might have on human physiology.

This method was utilized in this project in an effort to determine whether changes in

Page 126: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

126

the activity of ZnT8 in the α-cell could contribute to the effects of ZnT8 risk alleles in

humans.

6.2 ZnT8 in the pancreatic α-cell

In agreement with earlier findings [222] we showed that although ZnT8 is highly

expressed in the α-cell, colocalization with glucagon is incomplete even more so than

in the β-cell. Thus, ZnT8 appears to reside in cellular locations other than glucagon

granules (Figure 3.1). Although the identity of such structures has not been explored

here a possibility is that they may at least partly be immature granules. Future

studies using higher resolution techniques (including super-resolution techniques

such as PALM/STORM) or immuno-electron microscopy will be needed to provide a

definitive answer.

To address the question regarding the role of ZnT8 in the α-cell and glucagon

secretion we began by generating a mouse that lacked ZnT8 specifically in the α-cell

only. Reports from other studies using the same PPG.Cre line that was used here

showed a range of recombination levels from 13% [298] to 45% [300] so it was

important to determine the extent of deletion in our mouse. FACS analysis revealed

that the recombination rate was quite low; with an overall ~15% reduction in ZnT8 in

the entire α-cell population. However, no deletion was observed in the β-cell

population demonstrating the selectivity of this approach.

Thus the observed effects may have been dampened compared to having had

complete or near complete ablation of ZnT8. Additionally, this could mean that any

subtle changes occurring due to the absence of ZnT8 might not have been detected

due to the low level of deletion achieved.

Page 127: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

127

αZnT8 KO mice displayed normal glucose tolerance as assessed by glucose tolerance

tests, and normal insulin sensitivity. There were no changes in fasting blood glucose

or fasting plasma glucagon levels. However, there was an effect on the release of

glucagon under low glucose conditions. Deletion of ZnT8 caused an increase in

glucagon secretion at hypoglycaemic conditions both in vitro and in vivo. These

changes in glucagon release cannot be accounted for by increased glucagon content

in KO mice since total pancreatic glucagon measured by an immunoassay was

unchanged as were glucagon mRNA levels. An increase in α-cell area of ~15% was

observed in the KO pancreas and this could be contributing to the enhanced

glucagon release but the increase was too small to account for the 2 fold change

detected in glucagon secretion by KO islets. No changes were observed in Ca2+

dynamics as peak amplitude and frequency of oscillations were comparable between

the two cell populations, WT and KO. Thus, the mechanism by which ZnT8 influences

glucagon release is Ca2+ independent.

We showed a reduction in granular Zn2+ content in KO cells suggesting that ZnT8 is

involved in the accumulation of Zn2+ in glucagon granules. Based on the decreased

Zn2+ content [268] and the findings of a loss of Zn2+ co-release with insulin when

ZnT8 is deleted in the β-cell [270], an elimination of Zn2+ from granules upon

exocytosis is to be expected from α-cells following ZnT8 ablation. The impaired

release of Zn2+ from α-cells could be involved in a mechanism via which ZnT8

suppresses glucagon release at low glucose concentrations. The likely reduction in

local Zn2+ levels could also be affecting secretion.

There have been several rodent studies looking into the effect of Zn2+ on glucagon

release, the vast majority demonstrating an inhibitory action of Zn2+. The Zn2+

chelator Ca2+-EDTA caused stimulation of secretion from the perfused rat pancreas

[84]. Incubation of isolated rat cells with Zn2+ diminished pyruvate and glucose

induced glucagon secretion through the reversible activation of KATP channels and

Page 128: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

128

subsequent reduction in electrical activity [121]. A similar study in the mouse

showed an inhibitory effect of Zn2+ on glucagon release on an α-cell line and isolated

islets with maximal inhibition at 10 µM Zn2+ [130]. This inhibitory action on glucagon

secretion was confirmed by another group experimenting with whole islets and FACS

sorted α-cell at low glucose concentrations [320]. However, the above effect of Zn2+

in the mouse cannot be attributed to KATP channel activation. Instead, it involved the

uptake of the ion by Ca2+ channels and a change in the redox state of the cell [130].

Thus the likely reduction in Zn2+ released from secretory granules could be alleviating

the suppression normally imposed by the ion on surrounding α-cells and could be

contributing to the enhanced glucagon secretion observed from KO islets (Figure

6.1).

Correspondingly, overexpression of ZnT8 led to a reciprocal decrease in glucagon

secretion from isolated islets at low glucose (Chapter 5). This further confirms the

importance of ZnT8 in maintaining normal α-cell responses during hypoglycaemia.

Research on mouse cell lines are in agreement with the findings in this study.

Overexpression of ZnT8 variants resulted in a reduction in glucagon secretion by 50%

while knockdown caused an increase in secretion by 70% [274]. The changes in

glucagon secretion were accompanied by corresponding changes in glucagon

content which was not observed in our study since total glucagon content was

unchanged.

Zn2+ is important in the β-cell for insulin crystallization and storage in the granules

[267]. No such role for Zn2+ in the α-cell has been described yet the presence of Zn2+

in glucagon granules has been confirmed by electron microscopy although the

staining is less intense that in insulin granules [302]. It is thus plausible that Zn2+ may

be involved in glucagon packaging within the granule. A less compact aggregation of

individual glucagon particles inside the granule could enable the release of the

peptide via the dilating fusion pore so that a bigger fraction of exocytotic events go

through to completion prior to pore closure [321].

Page 129: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

129

A

B

C

Figure 6.1 – Hypothetical model reconciling data on the effect of Zn2+ on glucagon

secretion and our mouse models. (A) Glucagon secretion occurring at normal levels from

WT α-cells with Zn2+ released from the granules during exocytosis. Zn2+ can enter

surrounding cells where it can influence glucagon secretion in an inhibitory manner. (B) α-

cells lacking ZnT8 have lower levels of Zn2+ in their secretory granules and subsequently

lesser amounts are released in the extracellular space. Less Zn2+ enters neighboring cells so

inhibition of glucagon secretion is lower thus more glucagon is released. (C) In ZnT8

overexpressing α-cells, secretory granules are expected to contain more Zn2+ meaning more

would be released in the extracellular space and more would enter the surrounding cells

imposing more inhibition on glucagon release. That would explain the lower levels of

glucagon secretion observed.

Page 130: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

130

The full effect of a complete loss of ZnT8 may not be apparent due to compensatory

gene expression changes possibly during development which have not yet been

identified. The obvious candidates are the other members of the ZnT family of

transporters. However, arguing against this mechanism gene expression analysis

showed the presence of mRNAs encoding several ZnTs in the mouse α-cell but there

was no evidence for a compensatory change in their expression levels in islets with

α-cell specific deletion of ZnT8. Similarly there was no compensatory increase in

expression following global [222] or β-cell specific ablation of ZnT8 [270, 272]. On

the other hand findings from knockdown studies in cell lines and KO mouse models

indicate that ZnT3 [322, 323] and ZnT7 [324] could be involved in the control of

glucose stimulated insulin secretion in a ZnT8 independent manner at least in mice.

ZnT7 deletion caused decreased insulin secretion and increased insulin resistance

leading to glucose intolerance [324]. Of note, however multiple studies form our

laboratory have failed to measure detectable levels of ZnT3 mRNA in mouse islets

[222, 325].

The observed enhancement in glucagon secretion was achieved when ZnT8 was

deleted from only a small proportion of α-cells. This indicates that the absence of

ZnT8 has a large effect at the level of individual α-cells or that the KO cells may have

a disproportionate role in regulating the activity of the rest of the α-cell population.

The data gathered from this study does not eliminate either possibility. It is unlikely

that this increase in total secretion is achieved by an elevation of release at the level

of individual α-cells lacking ZnT8 since they would have to secrete ~7 fold more to

achieve a 2-fold change in glucagon levels.

Another possibility is that α-cells are highly inter-connected (and therefore inter-

dependent) as reported by our laboratory to be the case for β-cells [326]. The islet

with its 3D structure is thought to be conducive to intercellular signaling for

coordination of insulin secretion. The distinct flow of information through

Page 131: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

131

interconnected cells causes an organized release of insulin responding to glucose or

other mediators.

6.3 ZnT8 and Zn2+ homeostasis in the α-cell

Cytosolic free Zn2+ measurements were performed using the molecular Zn2+ probe

FluoZin-3 and revealed significant reductions in cytosolic levels following ZnT8

ablation (Chapter 4). Considering that ZnT8 was only deleted in 50% of the RFP

population it is possible that the lowering of Zn2+ levels was underestimated. A

decrease in Zn2+ was also detected in the granules using ZinPyr-4 a molecular probe

that localizes to the granules. These effects of ZnT8 deletion are consistent with

observations in the β-cell [222, 268]. The reduction in granular Zn2+ levels is in

agreement with the proposed role of ZnT8 as an efflux transporter on granule

membranes. On the other hand, the decrease in cytosolic Zn2+ is less easy to explain

but may reflect a decrease in cytosolic Zn2+ uptake across the plasma membrane.

Consideration of the biology of another member of the ZnT family, ZnT5, suggests an

alternative possibility. ZnT5 is ubiquitously expressed and more abundant in islets

compared to other tissues. It has been suggested that it may have the ability for

bidirectional transport [311]. It seems to have discrete roles in Zn2+ homeostasis

since in addition to being expressed on secretory granules and the Golgi apparatus in

the β-cell, [247] it was found to be expressed on the plasma membrane where it is

thought to mediate Zn2+ influx instead of efflux [311]. ZnT8 immunostaining was

detected at the plasma membrane as well as secretory granules [222]. A similar role

in influx as ZnT5 at that location could explain the reduction in cytosolic Zn2+ levels

observed both in α-cells and β-cells following cell specific deletion of ZnT8.

Transfection of HeLa cells, which do not normally express ZnT8, with a ZnT8

construct led to an increase in cytosolic Zn2+ [213]. This finding is also consistent with

a role for ZnT8 in Zn2+ influx.

Page 132: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

132

6.4 ZnT8, SLC30A8 and T2D risk

The mechanism involved in driving the association between rs13266634 risk alleles

and T2D has previously been investigated. Nicolson et al [222] observed the

accumulation of Zn2+ in the cell using molecular probes following transient

expression of either the Arg or Trp variants in a β-cell line. Cells expressing the Trp

construct had a greater fluorescence change suggesting that this variant is a more

active transporter. How this change, i.e. increased cytosolic free Zn2+ is actually

beneficial to the β-cell is still unclear since the Trp variant is not linked to improved

basal or glucose stimulated insulin secretion from human islets [327]. In addition the

report by Flannick et al [221] that haploinsufficiency for SLC30A8 is associated with

significantly (>50%) reduced risk of T2D contradicts the finding of a slightly less

active ZnT8 causing an increase in T2D risk. Thus it appears that a more intricate

mechanism is accountable for the rs13266634 association to T2D and the role of

ZnT8 in the control of hormone release [328].

The complexity regarding the role of ZnT8 could at least partly involve the

interaction between the pancreas and the liver. Zn2+ co-secreted with insulin plays a

part in the regulation of hepatic hormone clearance [271] where is suppresses

uptake by inhibiting clathrin-dependent endocytosis.

Taking into consideration the effect of haploinsufficiency potential explanations are

that the Trp variant turns over faster than the Arg one or that it traffics to different

sub-cellular locations [223]. The latter seems likely since the biggest difference

between the Arg and Trp variants was observed during measurements of

cytoplasmic Zn2+ accumulation [222] suggesting ZnT8 can transport Zn2+

bidirectionally being involved in influx across the plasma membrane. A possibility is

that the protective effect of the Trp variant may arise from increased activity at the

plasma membrane although how this is protective is as yet unclear [223].

Page 133: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

133

A hypothetical model has been proposed by Davidson and colleagues [223] to bring

together data for rare and common SLC30A8 variants. Glucose stimulated insulin

secretion leads to an elevated production of reactive oxygen and nitrogen species

[329]. Taking into account the fact that Zn2+ plays a key part in moderating oxidative

stress [330] a theory that could connect the contrasting observations is that the Trp

and loss-of-functions variants work by briefly delivering more Zn2+ into the

cytoplasm during secretion (Figure 6.2). This increase in Zn2+ in turn might reduce

the deleterious effects of reactive oxygen species. The more active Trp variant at the

plasma membrane would increase cytosolic Zn2+ by an increased reuptake of Zn2+ co-

secreted with insulin from the extracellular space. In the case of the loss-of-function

variants, their presence at the granule membrane would similarly lead to an increase

in cytosolic Zn2+ through a reduction in transport of the ion in the granule lumen

[223]. This is a model suggested for the β-cells but something similar could be

occurring in the α-cell to explain the observations on T2D risk.

Although plausible there is a weakness in this scheme since it implies that a loss of

function variant would actually lead to an increase in cytosolic Zn2+ levels whereas in

KO studies the opposite has been observed. In the α-cell we showed a reduction in

cytosolic Zn2+ following ZnT8 ablation. Similarly β-cell specific deletion led to a

reduction in intracellular Zn2+ levels [222, 268]. Thus it appears that the relationship

between the ZnT8 variants and T2D risk is more complicated.

Page 134: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

134

Figure 6.2 – Hypothetical model reconciling data for rare and common SLC30A8 variants in

the β-cell. The speculation is that the rs13266634 allele encoding the 325 Trp variant and

the rare mutations leading to haploinsufficiency are protective due to prolonging the

transient elevation in cytosolic Zn2+ levels occurring during glucose stimulated insulin

secretion. This subsequently reduces the harmful effects of oxygen reactive species. In (A)

the 325 Trp variant increases cytosolic Zn2+ levels by enhanced reuptake of Zn2+ secreted

alongside insulin at the plasma membrane. In (B) rare loss of function mutations lead to an

increase in Zn2+ levels primarily by reduced uptake of Zn2+ into secretory granules. Other ZnT

family members could be compensating for the possible decrease in ZnT8 facilitated uptake

at the cell membrane. Adapted from Davidson et al [223].

A

B

Page 135: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

135

Considering that the activity of the Arg form of ZnT8 is 30% less than the Trp variant

[259], it is reasonable to assume that Zn2+ accumulation into granules would be

lowered by 15% and 30% in heterozygote and homozygote Arg allele carriers

respectively. Thus heterozygote loss-of-function would have a 50% reduction [328].

Data from mouse studies primarily suggest a harmful effect on glucose homeostasis

when ZnT8 is deleted in the β-cell yet it is likely that there are species-specific

differences in the role of ZnT8 and Zn2+ in the islet and in hepatic hormone clearance

[271]. It is likely that in humans the interplay between ZnT8 activity and disease risk

may involve a complex dose-response. Interplay between actions on α- and β-cells

may contribute to this complexity.

In this case even small reductions in ZnT8 activity, for example the Arg325 variant,

could be deleterious while in the case of the loss-of-function variant a more sizeable

reduction could be protective in a situation where enhanced insulin secretion offsets

the harmful effect of elevated clearance [328]. Conversely in the mouse there may

be a less complex relationship where elevated hepatic clearance as a result of falling

ZnT8 levels leads to increasingly impaired insulin action.

More recent studies seek to determine not just which loci are associated with T2D

but how such loci are associated with glycaemia and disease pathogenesis. Co-

expression analysis demonstrated that the expression of many genes correlated

strongly with glucagon. One of these genes was SLC30A8. The lower the ZnT8

expression was in human islets, the lower the expression of glucagon [331]. Such a

finding suggests that ZnT8 may be involved in the control of glucagon expression or

vice versa. This is not something we however detected in the mouse. Although the

KO islets expressed less ZnT8 the expression of glucagon did not differ significantly

when compared to WT islets.

Page 136: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

136

In addition, SLC30A8 was found to be a target for the transcription factor 7-like 2

(TCF7L2) by our laboratory [332], a finding later confirmed by others [331]. Variants

of this transcription factor are associated with increased risk of T2D, having the

highest effect on disease risk out of all the T2D variants uncovered so far by GWAS

[207]. It is thus possible that ZnT8 is involved in a regulatory network with other

diabetes-associated genes [333] so that specific variants of such genes may regulate

α- and β-cells and thus have an effect on insulin and glucagon levels and

consequently T2D risk.

6.5 Final conclusions

By creating tissue specific KO mice where ZnT8 is deleted specifically in α-cells, and

transgenic animals overexpressing this transporter inducibly in these cells, we

demonstrate that the Zn2+ transporter regulates normal glucagon secretion during

hypoglycaemia. Additionally we show that ZnT8 is important for intracellular Zn2+

homeostasis affecting both cytosolic and granular free Zn2+ levels. Changes in ZnT8

expression or activity in the α-cell may thus contribute to diabetes risk in SLC30A8

risk carriers. ZnT8 might therefore provide a suitable target for pharmacotherapeutic

interventions acting to modulate both insulin and glucagon secretion in the context

of both T2D and glucagon secretion during hypoglycaemia in T1D.

Page 137: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

137

6.6 Future directions

Future studies are needed to provide further insight into the role of ZnT8 in the α-

cell and to complement the findings from this study.

One of the major effects ZnT8 had on the β-cell was on the morphology of secretory

granules. It would be possible to investigate this in the α-cell by staining for glucagon

in order to identify them during electron microscopy.

RNA seq analysis of RFP+ cells from WT and KO islets would give information into

whether ZnT8 deletion causes any specific changes to the transcriptome in the α-

cells and whether any such changes could account for the observed differences in

glucagon secretion.

To assess the theory of Zn2+ contributing to the observed effects of deletion of ZnT8

on glucagon secretion we could use the Zn2+ chelator TPEN to check if lowering Zn2+

levels in the surrounding of the cells could further enhance the exaggerated

glucagon response in low glucose concentrations. Alternatively we could use the Zn2+

ionophore pyrithione to evaluate whether increased entry of the ion into the cell

dampens the elevated glucagon levels observed during hypoglycaemia in ZnT8 KO

mice.

Crossing the ZnT8TgGlurtTA mice with a mouse containing a fluorescent reporter

under Tet On control would allow its expression specifically in the α-cell allowing

experiments at the single cell level as those performed in the ZnT8 KO mice. It would

be interesting to see whether the ZnT8 overexpressing α-cells have a reciprocal

change in intracellular Zn2+ levels.

Page 138: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

138

Chapter 7

References

Page 139: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

139

1. Shih, H.P., A. Wang, and M. Sander, Pancreas organogenesis: from lineage determination to morphogenesis. Annu Rev Cell Dev Biol, 2013. 29: p. 81-105.

2. Scheele, G.A., Biosynthesis, segregation, and secretion of exportable proteins by the exocrine pancreas. Am J Physiol, 1980. 238(6): p. G467-77.

3. Islam, S., The islets of Langerhans. Advances in exprimental medicine and biology. Vol. 654. 2010, New York: Springer.

4. Herrera, P.L., et al., Embryogenesis of the murine endocrine pancreas; early expression of pancreatic polypeptide gene. Development, 1991. 113(4): p. 1257-65.

5. Teitelman, G., et al., Precursor cells of mouse endocrine pancreas coexpress insulin, glucagon and the neuronal proteins tyrosine hydroxylase and neuropeptide Y, but not pancreatic polypeptide. Development, 1993. 118(4): p. 1031-9.

6. Gromada, J., I. Franklin, and C.B. Wollheim, Alpha-cells of the endocrine pancreas: 35 years of research but the enigma remains. Endocr Rev, 2007. 28(1): p. 84-116.

7. Rutter, G.A., Nutrient-secretion coupling in the pancreatic islet beta-cell: recent advances. Mol Aspects Med, 2001. 22(6): p. 247-84.

8. Wierup, N., et al., The ghrelin cell: a novel developmentally regulated islet cell in the human pancreas. Regul Pept, 2002. 107(1-3): p. 63-9.

9. Cabrera, O., et al., The unique cytoarchitecture of human pancreatic islets has implications for islet cell function. Proc Natl Acad Sci U S A, 2006. 103(7): p. 2334-9.

10. Brissova, M., et al., Assessment of human pancreatic islet architecture and composition by laser scanning confocal microscopy. J Histochem Cytochem, 2005. 53(9): p. 1087-97.

11. Patel, Y.C., M. Amherdt, and L. Orci, Quantitative electron microscopic autoradiography of insulin, glucagon, and somatostatin binding sites on islets. Science, 1982. 217(4565): p. 1155-6.

12. Olsson, R. and P.-O. Carlsson, The pancreatic islet endothelial cell: Emerging roles in islet function and disease. The International Journal of Biochemistry & Cell Biology, 2006. 38(4): p. 492-497.

13. Henderson, J.R. and M.C. Moss, A MORPHOMETRIC STUDY OF THE ENDOCRINE AND EXOCRINE CAPILLARIES OF THE PANCREAS. Experimental Physiology, 1985. 70(3): p. 347-356.

14. Jansson, L. and C. Hellerstrom, Stimulation by glucose of the blood flow to the pancreatic islets of the rat. Diabetologia, 1983. 25(1): p. 45-50.

15. Brunicardi, F.C., et al., Microcirculation of the Islets of Langerhans: Long Beach Veterans Administration Regional Medical Education Center Symposium. Diabetes, 1996. 45(4): p. 385-392.

16. Cegrell, L., ADRENERGIC NERVES AND MONOAMINE-CONTAINING CELLS IN THE MAMMALIAN ENDOCRINE PANCREAS:A COMPARATIVE STUDY. Acta Physiologica Scandinavica, 1968. 73: p. 17-23.

17. Ahren, B., et al., Galanin is co-localized with noradrenaline and neuropeptide Y in dog pancreas and celiac ganglion. Cell Tissue Res, 1990. 261(1): p. 49-58.

18. Pettersson, M., et al., Neuropeptide Y: intrapancreatic neuronal localization and effects on insulin secretion in the mouse. Cell Tissue Res, 1987. 248(1): p. 43-8.

19. Vieira, E., Y.J. Liu, and E. Gylfe, Involvement of alpha1 and beta-adrenoceptors in adrenaline stimulation of the glucagon-secreting mouse alpha-cell. Naunyn Schmiedebergs Arch Pharmacol, 2004. 369(2): p. 179-83.

20. Porte, D., Jr. and R.H. Williams, Inhibition of insulin release by norepinephrine in man. Science, 1966. 152(3726): p. 1248-50.

21. Holst, J.J., et al., Autonomic nervous control of pancreatic somatostatin secretion. Am J Physiol, 1983. 245(6): p. E542-8.

Page 140: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

140

22. Coupland, R.E., The innervation of pan creas of the rat, cat and rabbit as revealed by the cholinesterase technique. J Anat, 1958. 92(1): p. 143-9.

23. Bishop, A.E., et al., The location of VIP in the pancreas of man and rat. Diabetologia, 1980. 18(1): p. 73-8.

24. Ahren, B. and G.J. Taborsky, Jr., The mechanism of vagal nerve stimulation of glucagon and insulin secretion in the dog. Endocrinology, 1986. 118(4): p. 1551-7.

25. Ahren, B., T.L. Paquette, and G.J. Taborsky, Jr., Effect and mechanism of vagal nerve stimulation on somatostatin secretion in dogs. Am J Physiol, 1986. 250(2 Pt 1): p. E212-7.

26. Sorenson, R.L., D.G. Garry, and T.C. Brelje, Structural and functional considerations of GABA in islets of Langerhans. Beta-cells and nerves. Diabetes, 1991. 40(11): p. 1365-74.

27. Ekblad, E., P. Alm, and F. Sundler, Distribution, origin and projections of nitric oxide synthase-containing neurons in gut and pancreas. Neuroscience, 1994. 63(1): p. 233-48.

28. Rehfeld, J.F., et al., Neural regulation of pancreatic hormone secretion by the C-terminal tetrapeptide of CCK. Nature, 1980. 284(5751): p. 33-8.

29. Coutinho-Silva, R., et al., Changes in expression of P2 receptors in rat and mouse pancreas during development and ageing. Cell Tissue Res, 2001. 306(3): p. 373-83.

30. Hayashi, M., et al., Differentiation-associated Na+-dependent inorganic phosphate cotransporter (DNPI) is a vesicular glutamate transporter in endocrine glutamatergic systems. J Biol Chem, 2001. 276(46): p. 43400-6.

31. Zern, R.T., J.L. Bird, and J.M. Feldman, Effect of increased pancreatic islet norepinephrine, dopamine and serotonin concentration on insulin secretion in the golden hamster. Diabetologia, 1980. 18(4): p. 341-6.

32. Rodriguez-Diaz, R., et al., Alpha cells secrete acetylcholine as a non-neuronal paracrine signal priming beta cell function in humans. Nat Med, 2011. 17(7): p. 888-92.

33. Murlin, J.R., et al., AQUEOUS EXTRACTS OF PANCREAS: I. INFLUENCE ON THE CARBOHYDRATE METABOLISM OF DEPANCREATIZED ANIMALS. Journal of Biological Chemistry, 1923. 56(1): p. 253-296.

34. Sutherland, E.W. and C. de Duve, ORIGIN AND DISTRIBUTION OF THE HYPERGLYCEMIC-GLYCOGENOLYTIC FACTOR OF THE PANCREAS. Journal of Biological Chemistry, 1948. 175(2): p. 663-674.

35. Foa, P.P., H.R. Weinstein, and J.A. Smith, Secretion of insulin and of a hyperglycemic substance studied by means of pancreatic-femoral cross-circulation experiments. Am J Physiol, 1949. 157(2): p. 197-204.

36. Staub, A., L. Sinn, and O.K. Behrens, Purification and crystallization of glucagon. J Biol Chem, 1955. 214(2): p. 619-32.

37. Bromer, W.W., et al., The Amino Acid Sequence of Glucagon. I. Amino Acid Composition and Terminal Amino Acid Analyses. Journal of the American Chemical Society, 1957. 79(11): p. 2794-2798.

38. Unger, R.H., et al., GLUCAGON ANTIBODIES AND AN IMMUNOASSAY FOR GLUCAGON*. The Journal of Clinical Investigation, 1961. 40(7): p. 1280-1289.

39. Baum, J., et al., Localization of glucagon in the alpha cells in the pancreatic islet by immunofluorescent technics. Diabetes, 1962. 11: p. 371-4.

40. Offield, M.F., et al., PDX-1 is required for pancreatic outgrowth and differentiation of the rostral duodenum. Development, 1996. 122(3): p. 983-95.

41. Gradwohl, G., et al., neurogenin3 is required for the development of the four endocrine cell lineages of the pancreas. Proc Natl Acad Sci U S A, 2000. 97(4): p. 1607-11.

Page 141: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

141

42. Naya, F.J., et al., Diabetes, defective pancreatic morphogenesis, and abnormal enteroendocrine differentiation in BETA2/neuroD-deficient mice. Genes Dev, 1997. 11(18): p. 2323-34.

43. Collombat, P., et al., Opposing actions of Arx and Pax4 in endocrine pancreas development. Genes Dev, 2003. 17(20): p. 2591-603.

44. Hussain, M.A., C.P. Miller, and J.F. Habener, Brn-4 transcription factor expression targeted to the early developing mouse pancreas induces ectopic glucagon gene expression in insulin-producing beta cells. J Biol Chem, 2002. 277(18): p. 16028-32.

45. Artner, I., et al., MafB: an activator of the glucagon gene expressed in developing islet alpha- and beta-cells. Diabetes, 2006. 55(2): p. 297-304.

46. St-Onge, L., et al., Pax6 is required for differentiation of glucagon-producing alpha-cells in mouse pancreas. Nature, 1997. 387(6631): p. 406-9.

47. Bell, G.I., R.F. Santerre, and G.T. Mullenbach, Hamster preproglucagon contains the sequence of glucagon and two related peptides. Nature, 1983. 302(5910): p. 716-718.

48. Bell, G.I., et al., Exon duplication and divergence in the human preproglucagon gene. Nature, 1983. 304(5924): p. 368-71.

49. Heinrich, G., et al., Pre-proglucagon messenger ribonucleic acid: nucleotide and encoded amino acid sequences of the rat pancreatic complementary deoxyribonucleic acid. Endocrinology, 1984. 115(6): p. 2176-81.

50. Steiner, D.F., et al., The new enzymology of precursor processing endoproteases. J Biol Chem, 1992. 267(33): p. 23435-8.

51. Cordier-Bussat, M., C. Morel, and J. Philippe, Homologous DNA sequences and cellular factors are implicated in the control of glucagon and insulin gene expression. Molecular and Cellular Biology, 1995. 15(7): p. 3904-16.

52. Philippe, J., et al., Alpha-cell-specific expression of the glucagon gene is conferred to the glucagon promoter element by the interactions of DNA-binding proteins. Mol Cell Biol, 1988. 8(11): p. 4877-88.

53. Bosco, D., et al., Unique arrangement of alpha- and beta-cells in human islets of Langerhans. Diabetes, 2010. 59(5): p. 1202-10.

54. Philippe, J., C. Morel, and M. Cordier-Bussat, Islet-specific proteins interact with the insulin-response element of the glucagon gene. J Biol Chem, 1995. 270(7): p. 3039-45.

55. Ritz-Laser, B., et al., Pax-6 and Cdx-2/3 interact to activate glucagon gene expression on the G1 control element. J Biol Chem, 1999. 274(7): p. 4124-32.

56. Ritz-Laser, B., et al., The pancreatic beta-cell-specific transcription factor Pax-4 inhibits glucagon gene expression through Pax-6. Diabetologia, 2002. 45(1): p. 97-107.

57. Ritz-Laser, B., et al., Ectopic expression of the beta-cell specific transcription factor Pdx1 inhibits glucagon gene transcription. Diabetologia, 2003. 46(6): p. 810-21.

58. Glick, E., D. Leshkowitz, and M.D. Walker, Transcription factor BETA2 acts cooperatively with E2A and PDX1 to activate the insulin gene promoter. J Biol Chem, 2000. 275(3): p. 2199-204.

59. Philippe, J., Insulin regulation of the glucagon gene is mediated by an insulin-responsive DNA element. Proc Natl Acad Sci U S A, 1991. 88(16): p. 7224-7.

60. Grzeskowiak, R., et al., Insulin responsiveness of the glucagon gene conferred by interactions between proximal promoter and more distal enhancer-like elements involving the paired-domain transcription factor Pax6. J Biol Chem, 2000. 275(39): p. 30037-45.

61. Furuta, M., et al., Severe defect in proglucagon processing in islet A-cells of prohormone convertase 2 null mice. J Biol Chem, 2001. 276(29): p. 27197-202.

Page 142: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

142

62. Mayo, K.E., et al., International Union of Pharmacology. XXXV. The glucagon receptor family. Pharmacol Rev, 2003. 55(1): p. 167-94.

63. Quesada, I., et al., Physiology of the pancreatic alpha-cell and glucagon secretion: role in glucose homeostasis and diabetes. J Endocrinol, 2008. 199(1): p. 5-19.

64. Hansen, L.H., et al., The Gly40Ser mutation in the human glucagon receptor gene associated with NIDDM results in a receptor with reduced sensitivity to glucagon. Diabetes, 1996. 45(6): p. 725-30.

65. Huang, C.N., et al., Screening for the Gly40Ser mutation in the glucagon receptor gene among patients with type 2 diabetes or essential hypertension in Taiwan. Pancreas, 1999. 18(2): p. 151-5.

66. Parker, J.C., et al., Glycemic control in mice with targeted disruption of the glucagon receptor gene. Biochem Biophys Res Commun, 2002. 290(2): p. 839-43.

67. Frayn, K.N., Metabolic Regulation: A Human Perspective. 2nd ed. 2003, Oxford: Blackwell Science.

68. Unger, R.H. and L. Orci, Physiology and pathophysiology of glucagon. Physiol Rev, 1976. 56(4): p. 778-826.

69. Herzig, S., et al., CREB regulates hepatic gluconeogenesis through the coactivator PGC-1. Nature, 2001. 413(6852): p. 179-83.

70. Yoon, J.C., et al., Control of hepatic gluconeogenesis through the transcriptional coactivator PGC-1. Nature, 2001. 413(6852): p. 131-8.

71. Koo, S.H., et al., The CREB coactivator TORC2 is a key regulator of fasting glucose metabolism. Nature, 2005. 437(7062): p. 1109-11.

72. Kurland, I.J. and S.J. Pilkis, Covalent control of 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase: insights into autoregulation of a bifunctional enzyme. Protein Sci, 1995. 4(6): p. 1023-37.

73. Slavin, B.G., J.M. Ong, and P.A. Kern, Hormonal regulation of hormone-sensitive lipase activity and mRNA levels in isolated rat adipocytes. J Lipid Res, 1994. 35(9): p. 1535-41.

74. Ciudad, C., et al., Control of glycogen synthase phosphorylation in isolated rat hepatocytes by epinephrine, vasopressin and glucagon. Eur J Biochem, 1984. 142(3): p. 511-20.

75. Andersen, B., et al., Inhibition of glycogenolysis in primary rat hepatocytes by 1, 4-dideoxy-1,4-imino-D-arabinitol. Biochem J, 1999. 342 Pt 3: p. 545-50.

76. Cynober, L.A., Plasma amino acid levels with a note on membrane transport: characteristics, regulation, and metabolic significance. Nutrition, 2002. 18(9): p. 761-6.

77. Barg, S., et al., Tight coupling between electrical activity and exocytosis in mouse glucagon-secreting alpha-cells. Diabetes, 2000. 49(9): p. 1500-1510.

78. Pfeifer, C.R., et al., Quantitative analysis of mouse pancreatic islet architecture by serial block-face SEM. Journal of Structural Biology, 2015. 189(1): p. 44-52.

79. McTaggart, J.S., R.H. Clark, and F.M. Ashcroft, The role of the KATP channel in glucose homeostasis in health and disease: more than meets the islet. J Physiol, 2010. 588(Pt 17): p. 3201-9.

80. Rorsman, P. and B. Hellman, Voltage-activated currents in guinea pig pancreatic alpha 2 cells. Evidence for Ca2+-dependent action potentials. J Gen Physiol, 1988. 91(2): p. 223-42.

81. Olsen, H.L., et al., Glucose stimulates glucagon release in single rat alpha-cells by mechanisms that mirror the stimulus-secretion coupling in beta-cells. Endocrinology, 2005. 146(11): p. 4861-70.

82. Gopel, S.O., et al., Regulation of glucagon release in mouse -cells by KATP channels and inactivation of TTX-sensitive Na+ channels. J Physiol, 2000. 528(Pt 3): p. 509-20.

Page 143: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

143

83. Hope, K.M., et al., Regulation of alpha-cell function by the beta-cell in isolated human and rat islets deprived of glucose: the "switch-off" hypothesis. Diabetes, 2004. 53(6): p. 1488-95.

84. Ishihara, H., et al., Islet beta-cell secretion determines glucagon release from neighbouring alpha-cells. Nat Cell Biol, 2003. 5(4): p. 330-5.

85. MacDonald, P.E., et al., A K ATP channel-dependent pathway within alpha cells regulates glucagon release from both rodent and human islets of Langerhans. PLoS Biol, 2007. 5(6): p. e143.

86. Pipeleers, D.G., et al., Interplay of nutrients and hormones in the regulation of glucagon release. Endocrinology, 1985. 117(3): p. 817-23.

87. Johansson, H., E. Gylfe, and B. Hellman, Cyclic AMP raises cytoplasmic calcium in pancreatic alpha 2-cells by mobilizing calcium incorporated in response to glucose. Cell Calcium, 1989. 10(4): p. 205-11.

88. Ravier, M.A. and G.A. Rutter, Glucose or insulin, but not zinc ions, inhibit glucagon secretion from mouse pancreatic alpha-cells. Diabetes, 2005. 54(6): p. 1789-97.

89. Miki, T., et al., ATP-sensitive K+ channels in the hypothalamus are essential for the maintenance of glucose homeostasis. Nat Neurosci, 2001. 4(5): p. 507-12.

90. Sherck, S.M., et al., Pancreatic response to mild non-insulin-induced hypoglycemia does not involve extrinsic neural input. Diabetes, 2001. 50(11): p. 2487-96.

91. Shiota, C., et al., Impaired glucagon secretory responses in mice lacking the type 1 sulfonylurea receptor. Am J Physiol Endocrinol Metab, 2005. 289(4): p. E570-7.

92. Ramracheya, R., et al., Membrane potential-dependent inactivation of voltage-gated ion channels in alpha-cells inhibits glucagon secretion from human islets. Diabetes, 2010. 59(9): p. 2198-208.

93. Barg, S., Mechanisms of exocytosis in insulin-secreting B-cells and glucagon-secreting A-cells. Pharmacol Toxicol, 2003. 92(1): p. 3-13.

94. Gromada, J., et al., ATP-sensitive K+ channel-dependent regulation of glucagon release and electrical activity by glucose in wild-type and SUR1-/- mouse alpha-cells. Diabetes, 2004. 53 Suppl 3: p. S181-9.

95. Liu, Y.J., E. Vieira, and E. Gylfe, A store-operated mechanism determines the activity of the electrically excitable glucagon-secreting pancreatic alpha-cell. Cell Calcium, 2004. 35(4): p. 357-65.

96. Manning Fox, J.E., et al., Oscillatory membrane potential response to glucose in islet beta-cells: a comparison of islet-cell electrical activity in mouse and rat. Endocrinology, 2006. 147(10): p. 4655-63.

97. Vieira, E., A. Salehi, and E. Gylfe, Glucose inhibits glucagon secretion by a direct effect on mouse pancreatic alpha cells. Diabetologia, 2007. 50(2): p. 370-9.

98. Hahn, H.J., M. Ziegler, and E. Mohr, Inhibition of glucagon secretion by glucose and glyceraldehyde on isolated islets of Wistar rats. FEBS Lett, 1974. 49(1): p. 100-2.

99. Gerich, J.E., M.A. Charles, and G.M. Grodsky, Characterization of the effects of arginine and glucose on glucagon and insulin release from the perfused rat pancreas. J Clin Invest, 1974. 54(4): p. 833-41.

100. Zhang, Q., et al., Role of KATP channels in glucose-regulated glucagon secretion and impaired counterregulation in type 2 diabetes. Cell Metab, 2013. 18(6): p. 871-82.

101. Hermansen, K., Pancreatic D-cell recognition of D-glucose: studies with D-glucose, D-glyceraldehyde, dihydroxyacetone, D-mannoheptulose, D-fructose, D-galactose, and D-ribose. Diabetes, 1981. 30(3): p. 203-10.

102. Gerich, J.E., Glucose in the Control of Glucagon Secretion, in Glucagon II, P. Lefebvre, Editor. 1983, Springer Berlin Heidelberg. p. 3-18.

103. Heimberg, H., et al., The glucose sensor protein glucokinase is expressed in glucagon-producing alpha-cells. Proc Natl Acad Sci U S A, 1996. 93(14): p. 7036-41.

Page 144: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

144

104. Heimberg, H., et al., Differences in glucose transporter gene expression between rat pancreatic alpha- and beta-cells are correlated to differences in glucose transport but not in glucose utilization. J Biol Chem, 1995. 270(15): p. 8971-5.

105. Detimary, P., et al., The changes in adenine nucleotides measured in glucose-stimulated rodent islets occur in beta cells but not in alpha cells and are also observed in human islets. J Biol Chem, 1998. 273(51): p. 33905-8.

106. Gerich, J.E., M.A. Charles, and G.M. Grodsky, Regulation of pancreatic insulin and glucagon secretion. Annu Rev Physiol, 1976. 38: p. 353-88.

107. Edwards, J.C. and K.W. Taylor, Fatty acids and the release of glucagon from isolated guinea-pig islets of Langerhans incubated in vitro. Biochim Biophys Acta, 1970. 215(2): p. 310-5.

108. Andrews, S.S., S.A. Lopez, and W.G. Blackard, Effect of lipids on glucagon secretion in man. Metabolism, 1975. 24(1): p. 35-44.

109. Bollheimer, L.C., et al., Stimulatory short-term effects of free fatty acids on glucagon secretion at low to normal glucose concentrations. Metabolism, 2004. 53(11): p. 1443-8.

110. Hong, J., et al., The short-term effect of fatty acids on glucagon secretion is influenced by their chain length, spatial configuration, and degree of unsaturation: studies in vitro. Metabolism, 2005. 54(10): p. 1329-36.

111. Olofsson, C.S., et al., Palmitate stimulation of glucagon secretion in mouse pancreatic alpha-cells results from activation of L-type calcium channels and elevation of cytoplasmic calcium. Diabetes, 2004. 53(11): p. 2836-43.

112. Hong, J., et al., Fatty acid-induced effect on glucagon secretion is mediated via fatty acid oxidation. Diabetes Metab Res Rev, 2007. 23(3): p. 202-10.

113. Gremlich, S., et al., Fatty acids decrease IDX-1 expression in rat pancreatic islets and reduce GLUT2, glucokinase, insulin, and somatostatin levels. J Biol Chem, 1997. 272(48): p. 30261-9.

114. Huang, L., et al., Detection of exocytosis at individual pancreatic beta cells by amperometry at a chemically modified microelectrode. Proc Natl Acad Sci U S A, 1995. 92(21): p. 9608-12.

115. Unger, R.H., Glucagon physiology and pathophysiology in the light of new advances. Diabetologia, 1985. 28(8): p. 574-8.

116. Raskin, P., Y. Fujita, and R.H. Unger, Effect of insulin-glucose infusions on plasma glucagon levels in fasting diabetics and nondiabetics. J Clin Invest, 1975. 56(5): p. 1132-8.

117. Weir, G.C., et al., Glucagon secretion from the perfused pancreas of streptozotocin-treated rats. Diabetes, 1976. 25(4): p. 275-82.

118. Dunbar, J.C. and A. Brown, Glucagon secretion by dispersed alpha cell enriched islets from streptozotocin treated hamsters in perifusion. Horm Metab Res, 1984. 16(5): p. 221-5.

119. Ostenson, C.G. and C. Grebing, Evidence for metabolic regulation of pancreatic glucagon secretion by L-glutamine. Acta Endocrinol (Copenh), 1985. 108(3): p. 386-91.

120. Greenbaum, C.J., et al., Intra-islet insulin permits glucose to directly suppress pancreatic A cell function. J Clin Invest, 1991. 88(3): p. 767-73.

121. Franklin, I., et al., Beta-cell secretory products activate alpha-cell ATP-dependent potassium channels to inhibit glucagon release. Diabetes, 2005. 54(6): p. 1808-15.

122. Kisanuki, K., et al., Expression of insulin receptor on clonal pancreatic alpha cells and its possible role for insulin-stimulated negative regulation of glucagon secretion. Diabetologia, 1995. 38(4): p. 422-9.

Page 145: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

145

123. Maruyama, H., et al., Insulin within islets is a physiologic glucagon release inhibitor. J Clin Invest, 1984. 74(6): p. 2296-9.

124. Brunicardi, F.C., et al., Immunoneutralization of somatostatin, insulin, and glucagon causes alterations in islet cell secretion in the isolated perfused human pancreas. Pancreas, 2001. 23(3): p. 302-8.

125. Kawamori, D., et al., Insulin signaling in alpha cells modulates glucagon secretion in vivo. Cell Metab, 2009. 9(4): p. 350-61.

126. Leung, Y.M., et al., Insulin regulates islet alpha-cell function by reducing KATP channel sensitivity to adenosine 5'-triphosphate inhibition. Endocrinology, 2006. 147(5): p. 2155-62.

127. Xu, E., et al., Intra-islet insulin suppresses glucagon release via GABA-GABAA receptor system. Cell Metab, 2006. 3(1): p. 47-58.

128. Baker, E.N., et al., The structure of 2Zn pig insulin crystals at 1.5 A resolution. Philos Trans R Soc Lond B Biol Sci, 1988. 319(1195): p. 369-456.

129. Hutton, J.C., E.J. Penn, and M. Peshavaria, Low-molecular-weight constituents of isolated insulin-secretory granules. Bivalent cations, adenine nucleotides and inorganic phosphate. Biochem J, 1983. 210(2): p. 297-305.

130. Gyulkhandanyan, A.V., et al., Investigation of transport mechanisms and regulation of intracellular Zn2+ in pancreatic alpha-cells. J Biol Chem, 2008. 283(15): p. 10184-97.

131. Patel, Y.C., Somatostatin and its receptor family. Front Neuroendocrinol, 1999. 20(3): p. 157-98.

132. Kumar, U., et al., Subtype-selective expression of the five somatostatin receptors (hSSTR1-5) in human pancreatic islet cells: a quantitative double-label immunohistochemical analysis. Diabetes, 1999. 48(1): p. 77-85.

133. Hunyady, B., et al., Immunohistochemical localization of somatostatin receptor SST2A in the rat pancreas. Endocrinology, 1997. 138(7): p. 2632-5.

134. Strowski, M.Z., et al., Somatostatin inhibits insulin and glucagon secretion via two receptors subtypes: an in vitro study of pancreatic islets from somatostatin receptor 2 knockout mice. Endocrinology, 2000. 141(1): p. 111-7.

135. Luft, R., S. Efendic, and T. Hokfelt, Somatostatin--both hormone and neurotransmitter? Diabetologia, 1978. 14(1): p. 1-13.

136. Yoshimoto, Y., et al., Somatostatin induces hyperpolarization in pancreatic islet alpha cells by activating a G protein-gated K+ channel. FEBS Lett, 1999. 444(2-3): p. 265-9.

137. Gromada, J., et al., Gi2 proteins couple somatostatin receptors to low-conductance K+ channels in rat pancreatic alpha-cells. Pflugers Arch, 2001. 442(1): p. 19-26.

138. Gromada, J., et al., Somatostatin inhibits exocytosis in rat pancreatic alpha-cells by G(i2)-dependent activation of calcineurin and depriming of secretory granules. J Physiol, 2001. 535(Pt 2): p. 519-32.

139. Fehmann, H.C., M. Strowski, and B. Goke, Functional characterization of somatostatin receptors expressed on hamster glucagonoma cells. Am J Physiol, 1995. 268(1 Pt 1): p. E40-7.

140. Franklin, I.K. and C.B. Wollheim, GABA in the endocrine pancreas: its putative role as an islet cell paracrine-signalling molecule. J Gen Physiol, 2004. 123(3): p. 185-90.

141. Wendt, A., et al., Glucose inhibition of glucagon secretion from rat alpha-cells is mediated by GABA released from neighboring beta-cells. Diabetes, 2004. 53(4): p. 1038-45.

142. Rorsman, P., et al., Glucose-inhibition of glucagon secretion involves activation of GABAA-receptor chloride channels. Nature, 1989. 341(6239): p. 233-6.

Page 146: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

146

143. Gilon, P., et al., The influence of gamma-aminobutyric acid on hormone release by the mouse and rat endocrine pancreas. Endocrinology, 1991. 129(5): p. 2521-9.

144. Gaskins, H.R., et al., Glucose modulates gamma-aminobutyric acid release from the pancreatic beta TC6 cell line. J Biol Chem, 1995. 270(51): p. 30286-9.

145. Moriyama, Y. and M. Hayashi, Glutamate-mediated signaling in the islets of Langerhans: a thread entangled. Trends Pharmacol Sci, 2003. 24(10): p. 511-7.

146. Maechler, P. and C.B. Wollheim, Mitochondrial glutamate acts as a messenger in glucose-induced insulin exocytosis. Nature, 1999. 402(6762): p. 685-9.

147. Weaver, C.D., et al., Differential expression of glutamate receptor subtypes in rat pancreatic islets. J Biol Chem, 1996. 271(22): p. 12977-84.

148. Uehara, S., et al., Metabotropic glutamate receptor type 4 is involved in autoinhibitory cascade for glucagon secretion by alpha-cells of islet of Langerhans. Diabetes, 2004. 53(4): p. 998-1006.

149. Bertrand, G., et al., Glutamate stimulates glucagon secretion via an excitatory amino acid receptor of the AMPA subtype in rat pancreas. Eur J Pharmacol, 1993. 237(1): p. 45-50.

150. Hayashi, M., et al., Secretory granule-mediated co-secretion of L-glutamate and glucagon triggers glutamatergic signal transmission in islets of Langerhans. J Biol Chem, 2003. 278(3): p. 1966-74.

151. Tong, Q., R. Ouedraogo, and A.L. Kirchgessner, Localization and function of group III metabotropic glutamate receptors in rat pancreatic islets. Am J Physiol Endocrinol Metab, 2002. 282(6): p. E1324-33.

152. Drucker, D.J., Incretin action in the pancreas: potential promise, possible perils, and pathological pitfalls. Diabetes, 2013. 62(10): p. 3316-23.

153. De Marinis, Y.Z., et al., GLP-1 inhibits and adrenaline stimulates glucagon release by differential modulation of N- and L-type Ca2+ channel-dependent exocytosis. Cell Metab, 2010. 11(6): p. 543-53.

154. Ding, W.G., et al., Glucagon-like peptide I and glucose-dependent insulinotropic polypeptide stimulate Ca2+-induced secretion in rat alpha-cells by a protein kinase A-mediated mechanism. Diabetes, 1997. 46(5): p. 792-800.

155. Hazama, A., S. Hayashi, and Y. Okada, Cell surface measurements of ATP release from single pancreatic beta cells using a novel biosensor technique. Pflugers Arch, 1998. 437(1): p. 31-5.

156. Tuduri, E., et al., Inhibition of Ca2+ signaling and glucagon secretion in mouse pancreatic alpha-cells by extracellular ATP and purinergic receptors. Am J Physiol Endocrinol Metab, 2008. 294(5): p. E952-60.

157. Chapal, J., et al., Evidence for an A2-subtype adenosine receptor on pancreatic glucagon secreting cells. Br J Pharmacol, 1985. 86(3): p. 565-9.

158. Grapengiesser, E., et al., Glucose induces glucagon release pulses antisynchronous with insulin and sensitive to purinoceptor inhibition. Endocrinology, 2006. 147(7): p. 3472-7.

159. Stagner, J.I., E. Samols, and V. Marks, The anterograde and retrograde infusion of glucagon antibodies suggests that A cells are vascularly perfused before D cells within the rat islet. Diabetologia, 1989. 32(3): p. 203-6.

160. Ma, X., et al., Glucagon stimulates exocytosis in mouse and rat pancreatic alpha-cells by binding to glucagon receptors. Mol Endocrinol, 2005. 19(1): p. 198-212.

161. Vincent, M., et al., Abrogation of protein convertase 2 activity results in delayed islet cell differentiation and maturation, increased alpha-cell proliferation, and islet neogenesis. Endocrinology, 2003. 144(9): p. 4061-9.

Page 147: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

147

162. Gelling, R.W., et al., Lower blood glucose, hyperglucagonemia, and pancreatic alpha cell hyperplasia in glucagon receptor knockout mice. Proc Natl Acad Sci U S A, 2003. 100(3): p. 1438-43.

163. Liang, Y., et al., Reduction in glucagon receptor expression by an antisense oligonucleotide ameliorates diabetic syndrome in db/db mice. Diabetes, 2004. 53(2): p. 410-7.

164. Wierup, N., F. Sundler, and R.S. Heller, The islet ghrelin cell. J Mol Endocrinol, 2014. 52(1): p. R35-49.

165. Wierup, N., et al., Ghrelin is expressed in a novel endocrine cell type in developing rat islets and inhibits insulin secretion from INS-1 (832/13) cells. J Histochem Cytochem, 2004. 52(3): p. 301-10.

166. Salehi, A., et al., Effects of ghrelin on insulin and glucagon secretion: a study of isolated pancreatic islets and intact mice. Regul Pept, 2004. 118(3): p. 143-50.

167. Chuang, J.C., et al., Ghrelin directly stimulates glucagon secretion from pancreatic alpha-cells. Mol Endocrinol, 2011. 25(9): p. 1600-11.

168. Ahren, B., Autonomic regulation of islet hormone secretion--implications for health and disease. Diabetologia, 2000. 43(4): p. 393-410.

169. Evans, M.L., et al., Hypothalamic ATP-sensitive K + channels play a key role in sensing hypoglycemia and triggering counterregulatory epinephrine and glucagon responses. Diabetes, 2004. 53(10): p. 2542-51.

170. Song, Z., et al., Convergence of pre- and postsynaptic influences on glucosensing neurons in the ventromedial hypothalamic nucleus. Diabetes, 2001. 50(12): p. 2673-81.

171. McCrimmon, R.J., et al., Activation of ATP-sensitive K+ channels in the ventromedial hypothalamus amplifies counterregulatory hormone responses to hypoglycemia in normal and recurrently hypoglycemic rats. Diabetes, 2005. 54(11): p. 3169-74.

172. Walker, J.N., et al., Regulation of glucagon secretion by glucose: paracrine, intrinsic or both? Diabetes Obes Metab, 2011. 13 Suppl 1: p. 95-105.

173. Ahren, B. and I. Lundquist, Influences of gastro-intestinal polypeptides and glucose on glucagon secretion induced by cholinergic stimulation. Horm Metab Res, 1982. 14(10): p. 529-32.

174. Gromada, J., et al., Adrenaline stimulates glucagon secretion in pancreatic A-cells by increasing the Ca2+ current and the number of granules close to the L-type Ca2+ channels. J Gen Physiol, 1997. 110(3): p. 217-28.

175. Diabetes in the UK 2012 - Key statistics on diabetes. 2012; Available from: http://www.diabetes.org.uk/diabetes-in-the-uk-2012.

176. Unger, R.H. and L. Orci, The essential role of glucagon in the pathogenesis of diabetes mellitus. Lancet, 1975. 1(7897): p. 14-6.

177. Standards of medical care in diabetes--2008. Diabetes Care, 2008. 31 Suppl 1: p. S12-54.

178. Coppieters, K.T., et al., Demonstration of islet-autoreactive CD8 T cells in insulitic lesions from recent onset and long-term type 1 diabetes patients. J Exp Med, 2012. 209(1): p. 51-60.

179. Roep, B.O., The role of T-cells in the pathogenesis of Type 1 diabetes: from cause to cure. Diabetologia, 2003. 46(3): p. 305-21.

180. Noble, J.A., et al., The role of HLA class II genes in insulin-dependent diabetes mellitus: molecular analysis of 180 Caucasian, multiplex families. Am J Hum Genet, 1996. 59(5): p. 1134-48.

181. Pipeleers, D. and Z. Ling, Pancreatic beta cells in insulin-dependent diabetes. Diabetes/Metabolism Reviews, 1992. 8(3): p. 209-227.

Page 148: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

148

182. Vaxillaire, M. and P. Froguel, Monogenic diabetes in the young, pharmacogenetics and relevance to multifactorial forms of type 2 diabetes. Endocr Rev, 2008. 29(3): p. 254-64.

183. Bonadonna, R.C., Alterations of glucose metabolism in type 2 diabetes mellitus. An overview. Rev Endocr Metab Disord, 2004. 5(2): p. 89-97.

184. Kahn, S.E., The relative contributions of insulin resistance and beta-cell dysfunction to the pathophysiology of Type 2 diabetes. Diabetologia, 2003. 46(1): p. 3-19.

185. Beard, J.C., et al., Relationship of islet function to insulin action in human obesity. J Clin Endocrinol Metab, 1987. 65(1): p. 59-64.

186. D'Alessio, D., The role of dysregulated glucagon secretion in type 2 diabetes. Diabetes Obes Metab, 2011. 13 Suppl 1: p. 126-32.

187. Ledermann, H.M., Is maturity onset diabetes at young age (MODY) more common in Europe than previously assumed? Lancet, 1995. 345(8950): p. 648.

188. Fajans, S.S., G.I. Bell, and K.S. Polonsky, Molecular mechanisms and clinical pathophysiology of maturity-onset diabetes of the young. N Engl J Med, 2001. 345(13): p. 971-80.

189. Froguel, P., et al., Familial hyperglycemia due to mutations in glucokinase. Definition of a subtype of diabetes mellitus. N Engl J Med, 1993. 328(10): p. 697-702.

190. Yamagata, K., et al., Mutations in the hepatocyte nuclear factor-4alpha gene in maturity-onset diabetes of the young (MODY1). Nature, 1996. 384(6608): p. 458-60.

191. Vaxillaire, M., et al., A gene for maturity onset diabetes of the young (MODY) maps to chromosome 12q. Nat Genet, 1995. 9(4): p. 418-23.

192. Horikawa, Y., et al., Mutation in hepatocyte nuclear factor-1 beta gene (TCF2) associated with MODY. Nat Genet, 1997. 17(4): p. 384-5.

193. Stoffers, D.A., et al., Early-onset type-II diabetes mellitus (MODY4) linked to IPF1. Nat Genet, 1997. 17(2): p. 138-9.

194. Malecki, M.T., et al., Mutations in NEUROD1 are associated with the development of type 2 diabetes mellitus. Nat Genet, 1999. 23(3): p. 323-8.

195. Zimmet, P., K.G. Alberti, and J. Shaw, Global and societal implications of the diabetes epidemic. Nature, 2001. 414(6865): p. 782-7.

196. O'Rahilly, S., et al., Type II diabetes of early onset: a distinct clinical and genetic syndrome? Br Med J (Clin Res Ed), 1987. 294(6577): p. 923-8.

197. Knowler, W.C., et al., Diabetes mellitus in the Pima Indians: incidence, risk factors and pathogenesis. Diabetes Metab Rev, 1990. 6(1): p. 1-27.

198. Barnett, A.H., et al., Diabetes in identical twins. A study of 200 pairs. Diabetologia, 1981. 20(2): p. 87-93.

199. Billings, L.K. and J.C. Florez, The genetics of type 2 diabetes: what have we learned from GWAS? Ann N Y Acad Sci, 2010. 1212: p. 59-77.

200. Abecasis, G.R., et al., A map of human genome variation from population-scale sequencing. Nature, 2010. 467(7319): p. 1061-73.

201. Griffith, O.L., et al., ORegAnno: an open-access community-driven resource for regulatory annotation. Nucleic Acids Res, 2008. 36(Database issue): p. D107-13.

202. Altshuler, D.M., et al., Integrating common and rare genetic variation in diverse human populations. Nature, 2010. 467(7311): p. 52-8.

203. Sladek, R., et al., A genome-wide association study identifies novel risk loci for type 2 diabetes. Nature, 2007. 445(7130): p. 881-5.

204. Steinthorsdottir, V., et al., A variant in CDKAL1 influences insulin response and risk of type 2 diabetes. Nat Genet, 2007. 39(6): p. 770-5.

205. Zeggini, E., et al., Replication of genome-wide association signals in UK samples reveals risk loci for type 2 diabetes. Science, 2007. 316(5829): p. 1336-41.

Page 149: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

149

206. Grant, S.F., et al., Variant of transcription factor 7-like 2 (TCF7L2) gene confers risk of type 2 diabetes. Nat Genet, 2006. 38(3): p. 320-3.

207. Helgason, A., et al., Refining the impact of TCF7L2 gene variants on type 2 diabetes and adaptive evolution. Nat Genet, 2007. 39(2): p. 218-25.

208. Florez, J.C., et al., TCF7L2 polymorphisms and progression to diabetes in the Diabetes Prevention Program. N Engl J Med, 2006. 355(3): p. 241-50.

209. Marullo, L., J.S. El-Sayed Moustafa, and I. Prokopenko, Insights into the genetic susceptibility to type 2 diabetes from genome-wide association studies of glycaemic traits. Curr Diab Rep, 2014. 14(11): p. 551.

210. da Silva Xavier, G., et al., Animal models of GWAS-identified type 2 diabetes genes. J Diabetes Res, 2013. 2013: p. 906590.

211. Gaulton, K.J., et al., Genetic fine mapping and genomic annotation defines causal mechanisms at type 2 diabetes susceptibility loci. 2015. 47(12): p. 1415-25.

212. Rutter, G.A., Dorothy Hodgkin Lecture 2014. Understanding genes identified by genome-wide association studies for type 2 diabetes. Diabet Med, 2014. 31(12): p. 1480-7.

213. Chimienti, F., et al., Identification and cloning of a beta-cell-specific zinc transporter, ZnT-8, localized into insulin secretory granules. Diabetes, 2004. 53(9): p. 2330-7.

214. Ng, M.C., et al., Implication of genetic variants near TCF7L2, SLC30A8, HHEX, CDKAL1, CDKN2A/B, IGF2BP2, and FTO in type 2 diabetes and obesity in 6,719 Asians. Diabetes, 2008. 57(8): p. 2226-33.

215. Borowiec, M., et al., Mutations in the SLC30A8 gene are not a major cause of MODY or other forms of early-onset, autosomal dominant type 2 diabetes. Diabetologia, 2007. 50(10): p. 2224-6.

216. Wenzlau, J.M., et al., The cation efflux transporter ZnT8 (Slc30A8) is a major autoantigen in human type 1 diabetes. Proc Natl Acad Sci U S A, 2007. 104(43): p. 17040-5.

217. Staiger, H., et al., Polymorphisms within novel risk loci for type 2 diabetes determine beta-cell function. PLoS One, 2007. 2(9): p. e832.

218. Kirchhoff, K., et al., Polymorphisms in the TCF7L2, CDKAL1 and SLC30A8 genes are associated with impaired proinsulin conversion. Diabetologia, 2008. 51(4): p. 597-601.

219. Cauchi, S., et al., Analysis of novel risk loci for type 2 diabetes in a general French population: the D.E.S.I.R. study. J Mol Med (Berl), 2008. 86(3): p. 341-8.

220. Pascoe, L., et al., Common variants of the novel type 2 diabetes genes CDKAL1 and HHEX/IDE are associated with decreased pancreatic beta-cell function. Diabetes, 2007. 56(12): p. 3101-4.

221. Flannick, J., et al., Loss-of-function mutations in SLC30A8 protect against type 2 diabetes. 2014. 46(4): p. 357-63.

222. Nicolson, T.J., et al., Insulin storage and glucose homeostasis in mice null for the granule zinc transporter ZnT8 and studies of the type 2 diabetes-associated variants. Diabetes, 2009. 58(9): p. 2070-83.

223. Davidson, H.W., J.M. Wenzlau, and R.M. O'Brien, Zinc transporter 8 (ZnT8) and beta cell function. Trends Endocrinol Metab, 2014. 25(8): p. 415-24.

224. Muller, W.A., G.R. Faloona, and R.H. Unger, Hyperglucagonemia in diabetic ketoacidosis. Its prevalence and significance. Am J Med, 1973. 54(1): p. 52-7.

225. Baron, A.D., et al., Role of hyperglucagonemia in maintenance of increased rates of hepatic glucose output in type II diabetics. Diabetes, 1987. 36(3): p. 274-83.

226. Unger, R.H., et al., Studies of pancreatic alpha cell function in normal and diabetic subjects. J Clin Invest, 1970. 49(4): p. 837-48.

Page 150: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

150

227. Bolli, G.B., et al., Defective glucose counterregulation after subcutaneous insulin in noninsulin-dependent diabetes mellitus. Paradoxical suppression of glucose utilization and lack of compensatory increase in glucose production, roles of insulin resistance, abnormal neuroendocrine responses, and islet paracrine interactions. J Clin Invest, 1984. 73(6): p. 1532-41.

228. Meneilly, G.S., E. Cheung, and H. Tuokko, Counterregulatory hormone responses to hypoglycemia in the elderly patient with diabetes. Diabetes, 1994. 43(3): p. 403-10.

229. Liljenquist, J.E., et al., Evidence for an important role of glucagon in the regulation of hepatic glucose production in normal man. J Clin Invest, 1977. 59(2): p. 369-74.

230. Lee, Y., et al., Hyperglycemia in rodent models of type 2 diabetes requires insulin-resistant alpha cells. Proc Natl Acad Sci U S A, 2014. 111(36): p. 13217-22.

231. Jansen, J., et al., Disturbed zinc homeostasis in diabetic patients by in vitro and in vivo analysis of insulinomimetic activity of zinc. J Nutr Biochem, 2012. 23(11): p. 1458-66.

232. Rutter, G.A., et al., Intracellular zinc in insulin secretion and action: a determinant of diabetes risk? Proc Nutr Soc, 2015: p. 1-12.

233. Scott, D.A. and A.M. Fisher, THE INSULIN AND THE ZINC CONTENT OF NORMAL AND DIABETIC PANCREAS. J Clin Invest, 1938. 17(6): p. 725-8.

234. Quarterman, J., C.F. Mills, and W.R. Humphries, The reduced secretion of, and sensitivity to insulin in zinc-deficient rats. Biochem Biophys Res Commun, 1966. 25(3): p. 354-8.

235. Boquist, L., et al., [Insulin biosynthesis, storage and secretion. 8. Pancreatic-islet tissue and heavy metals--some ultrastructural and experimental observations]. Lakartidningen, 1968. 65(37): p. 3603-7.

236. Ripa, S. and R. Ripa, [Zinc and diabetes mellitus]. Minerva Med, 1995. 86(10): p. 415-21.

237. Gupta, R., et al., Oral zinc therapy in diabetic neuropathy. J Assoc Physicians India, 1998. 46(11): p. 939-42.

238. Begin-Heick, N., et al., Zinc supplementation attenuates insulin secretory activity in pancreatic islets of the ob/ob mouse. Diabetes, 1985. 34(2): p. 179-84.

239. Simon, S.F. and C.G. Taylor, Dietary zinc supplementation attenuates hyperglycemia in db/db mice. Exp Biol Med (Maywood), 2001. 226(1): p. 43-51.

240. Chasapis, C.T., et al., Zinc and human health: an update. Arch Toxicol, 2012. 86(4): p. 521-34.

241. Andreini, C., et al., Counting the zinc-proteins encoded in the human genome. J Proteome Res, 2006. 5(1): p. 196-201.

242. Maret, W. and A. Krezel, Cellular zinc and redox buffering capacity of metallothionein/thionein in health and disease. Mol Med, 2007. 13(7-8): p. 371-5.

243. Maret, W., New perspectives of zinc coordination environments in proteins. J Inorg Biochem, 2012. 111: p. 110-6.

244. Yang, L., et al., Polymorphisms in metallothionein-1 and -2 genes associated with the risk of type 2 diabetes mellitus and its complications. Am J Physiol Endocrinol Metab, 2008. 294(5): p. E987-92.

245. Eide, D.J., Zinc transporters and the cellular trafficking of zinc. Biochim Biophys Acta, 2006. 1763(7): p. 711-22.

246. Palmiter, R.D. and L. Huang, Efflux and compartmentalization of zinc by members of the SLC30 family of solute carriers. Pflugers Arch, 2004. 447(5): p. 744-51.

247. Kambe, T., et al., Cloning and characterization of a novel mammalian zinc transporter, zinc transporter 5, abundantly expressed in pancreatic beta cells. J Biol Chem, 2002. 277(21): p. 19049-55.

Page 151: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

151

248. Huang, L., C.P. Kirschke, and J. Gitschier, Functional characterization of a novel mammalian zinc transporter, ZnT6. J Biol Chem, 2002. 277(29): p. 26389-95.

249. Fukunaka, A., et al., Demonstration and characterization of the heterodimerization of ZnT5 and ZnT6 in the early secretory pathway. J Biol Chem, 2009. 284(45): p. 30798-806.

250. Cousins, R.J., J.P. Liuzzi, and L.A. Lichten, Mammalian zinc transport, trafficking, and signals. J Biol Chem, 2006. 281(34): p. 24085-9.

251. Myers, S.A., A. Nield, and M. Myers, Zinc transporters, mechanisms of action and therapeutic utility: implications for type 2 diabetes mellitus. J Nutr Metab, 2012. 2012: p. 173712.

252. Wijesekara, N., F. Chimienti, and M.B. Wheeler, Zinc, a regulator of islet function and glucose homeostasis. Diabetes Obes Metab, 2009. 11 Suppl 4: p. 202-14.

253. Seve, M., et al., In silico identification and expression of SLC30 family genes: an expressed sequence tag data mining strategy for the characterization of zinc transporters' tissue expression. BMC Genomics, 2004. 5(1): p. 32.

254. Smidt, K., et al., Zinc-transporter genes in human visceral and subcutaneous adipocytes: lean versus obese. Mol Cell Endocrinol, 2007. 264(1-2): p. 68-73.

255. Overbeck, S., et al., Intracellular zinc homeostasis in leukocyte subsets is regulated by different expression of zinc exporters ZnT-1 to ZnT-9. J Leukoc Biol, 2008. 83(2): p. 368-80.

256. Murgia, C., et al., Diabetes-linked zinc transporter ZnT8 is a homodimeric protein expressed by distinct rodent endocrine cell types in the pancreas and other glands. Nutr Metab Cardiovasc Dis, 2009. 19(6): p. 431-9.

257. Lu, M. and D. Fu, Structure of the zinc transporter YiiP. Science, 2007. 317(5845): p. 1746-8.

258. Cherezov, V., et al., Insights into the mode of action of a putative zinc transporter CzrB in Thermus thermophilus. Structure, 2008. 16(9): p. 1378-88.

259. Kim, I., et al., A low-risk ZnT-8 allele (W325) for post-transplantation diabetes mellitus is protective against cyclosporin A-induced impairment of insulin secretion. Pharmacogenomics J, 2011. 11(3): p. 191-8.

260. Weijers, R.N., 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, 2010. 2(1): p. 33.

261. Pound, L.D., et al., The pancreatic islet beta-cell-enriched transcription factor Pdx-1 regulates Slc30a8 gene transcription through an intronic enhancer. Biochem J, 2011. 433(1): p. 95-105.

262. Pound, L.D., et al., Characterization of the human SLC30A8 promoter and intronic enhancer. J Mol Endocrinol, 2011. 47(3): p. 251-9.

263. Fu, Y., et al., Down-regulation of ZnT8 expression in INS-1 rat pancreatic beta cells reduces insulin content and glucose-inducible insulin secretion. PLoS One, 2009. 4(5): p. e5679.

264. Egefjord, L., et al., Zinc transporter gene expression is regulated by pro-inflammatory cytokines: a potential role for zinc transporters in beta-cell apoptosis? BMC Endocr Disord, 2009. 9: p. 7.

265. Eizirik, D.L., et al., The human pancreatic islet transcriptome: expression of candidate genes for type 1 diabetes and the impact of pro-inflammatory cytokines. PLoS Genet, 2012. 8(3): p. e1002552.

266. Dunn, M.F., Zinc-ligand interactions modulate assembly and stability of the insulin hexamer -- a review. Biometals, 2005. 18(4): p. 295-303.

267. Emdin, S.O., et al., Role of zinc in insulin biosynthesis. Some possible zinc-insulin interactions in the pancreatic B-cell. Diabetologia, 1980. 19(3): p. 174-82.

Page 152: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

152

268. Pound, L.D., et al., Deletion of the mouse Slc30a8 gene encoding zinc transporter-8 results in impaired insulin secretion. Biochem J, 2009. 421(3): p. 371-6.

269. Pound, L.D., et al., The physiological effects of deleting the mouse SLC30A8 gene encoding zinc transporter-8 are influenced by gender and genetic background. PLoS One, 2012. 7(7): p. e40972.

270. Lemaire, K., et al., Insulin crystallization depends on zinc transporter ZnT8 expression, but is not required for normal glucose homeostasis in mice. Proc Natl Acad Sci U S A, 2009. 106(35): p. 14872-7.

271. Tamaki, M., et al., The diabetes-susceptible gene SLC30A8/ZnT8 regulates hepatic insulin clearance. J Clin Invest, 2013. 123(10): p. 4513-24.

272. Wijesekara, N., et al., Beta cell-specific Znt8 deletion in mice causes marked defects in insulin processing, crystallisation and secretion. Diabetologia, 2010. 53(8): p. 1656-68.

273. Hardy, A.B., et al., Regulation of glucagon secretion by zinc: lessons from the beta cell-specific Znt8 knockout mouse model. Diabetes Obes Metab, 2011. 13 Suppl 1: p. 112-7.

274. Souza SC, Q.L., Inouye KE, Roix JJ, Chen H, Zinc transporter ZnT-8 regulates insulin and glucagon secretion in Min6 and alphaTC1-9 pancreatic cell lines. Diabetologia, 2008. 51 Suppl 1: p. 501.

275. Sorensen, H., et al., Immunoneutralization of endogenous glucagon reduces hepatic glucose output and improves long-term glycemic control in diabetic ob/ob mice. Diabetes, 2006. 55(10): p. 2843-8.

276. DeFronzo, R.A., et al., Novel Agents for the Treatment of Type 2 Diabetes. Diabetes Spectrum, 2014. 27(2): p. 100-112.

277. Longuet, C., et al., The glucagon receptor is required for the adaptive metabolic response to fasting. Cell metabolism, 2008. 8(5): p. 359-371.

278. Mu, J., et al., Chronic treatment with a glucagon receptor antagonist lowers glucose and moderately raises circulating glucagon and glucagon-like peptide 1 without severe alpha cell hypertrophy in diet-induced obese mice. Diabetologia, 2011. 54(9): p. 2381-91.

279. Engel SS, R.M., Xu L, Andryuk PJ, Davies MJ, Kaufman KD, Goldstein BJ, Glycemic and lipid effects of the short-acting glucagon receptor antagonist MK-3577 in patients with type 2 diabetes. Diabetes, 2012. 61(Suppl 1:A266).

280. Herrera, P.L., Adult insulin- and glucagon-producing cells differentiate from two independent cell lineages. Development, 2000. 127(11): p. 2317-22.

281. Solomou, A., et al., The Zinc Transporter Slc30a8/ZnT8 Is Required in a Subpopulation of Pancreatic alpha-Cells for Hypoglycemia-induced Glucagon Secretion. J Biol Chem, 2015. 290(35): p. 21432-42.

282. Schneider, C.A., W.S. Rasband, and K.W. Eliceiri, NIH Image to ImageJ: 25 years of image analysis. Nat Meth, 2012. 9(7): p. 671-675.

283. Burdette, S.C., et al., ZP4, an improved neuronal Zn2+ sensor of the Zinpyr family. J Am Chem Soc, 2003. 125(7): p. 1778-87.

284. Burgess, A., et al., Loss of human Greatwall results in G2 arrest and multiple mitotic defects due to deregulation of the cyclin B-Cdc2/PP2A balance. Proc Natl Acad Sci U S A, 2010. 107(28): p. 12564-9.

285. Powers, A.C., et al., Proglucagon processing similar to normal islets in pancreatic alpha-like cell line derived from transgenic mouse tumor. Diabetes, 1990. 39(4): p. 406-14.

286. Hohmeier, H.E., et al., Isolation of INS-1-derived cell lines with robust ATP-sensitive K+ channel-dependent and -independent glucose-stimulated insulin secretion. Diabetes, 2000. 49(3): p. 424-430.

Page 153: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

153

287. Graham, F.L., et al., Characteristics of a human cell line transformed by DNA from human adenovirus type 5. J Gen Virol, 1977. 36(1): p. 59-74.

288. Vinkenborg, J.L., et al., Genetically encoded FRET sensors to monitor intracellular Zn2+ homeostasis. Nat Methods, 2009. 6(10): p. 737-40.

289. Kahn, S.E., et al., The beta cell lesion in type 2 diabetes: there has to be a primary functional abnormality. Diabetologia, 2009. 52(6): p. 1003-12.

290. Gerich, J.E., et al., Lack of glucagon response to hypoglycemia in diabetes: evidence for an intrinsic pancreatic alpha cell defect. Science, 1973. 182(4108): p. 171-3.

291. Scott, L.J., et al., A genome-wide association study of type 2 diabetes in Finns detects multiple susceptibility variants. Science, 2007. 316(5829): p. 1341-5.

292. Voight, B.F., et al., Twelve type 2 diabetes susceptibility loci identified through large-scale association analysis. Nat Genet, 2010. 42(7): p. 579-89.

293. Luche, H., et al., Faithful activation of an extra-bright red fluorescent protein in "knock-in" Cre-reporter mice ideally suited for lineage tracing studies. Eur J Immunol, 2007. 37(1): p. 43-53.

294. Parker, H.E., et al., Predominant role of active versus facilitative glucose transport for glucagon-like peptide-1 secretion. Diabetologia, 2012. 55(9): p. 2445-55.

295. Reaven, E.P., D.L. Curry, and G.M. Reaven, Effect of age and sex on rat endocrine pancreas. Diabetes, 1987. 36(12): p. 1397-400.

296. Orci, L., et al., Localization of the pancreatic beta cell glucose transporter to specific plasma membrane domains. Science, 1989. 245(4915): p. 295-7.

297. Quoix, N., et al., The GluCre-ROSA26EYFP mouse: a new model for easy identification of living pancreatic alpha-cells. FEBS Lett, 2007. 581(22): p. 4235-40.

298. Lu, J., et al., Alpha cell-specific Men1 ablation triggers the transdifferentiation of glucagon-expressing cells and insulinoma development. Gastroenterology, 2010. 138(5): p. 1954-65.

299. Wilcox, C.L., et al., Pancreatic alpha-cell specific deletion of mouse Arx leads to alpha-cell identity loss. PLoS One, 2013. 8(6): p. e66214.

300. Tuduri, E., et al., Partial ablation of leptin signaling in mouse pancreatic alpha-cells does not alter either glucose or lipid homeostasis. Am J Physiol Endocrinol Metab, 2014. 306(7): p. E748-55.

301. Vallee, B.L. and K.H. Falchuk, The biochemical basis of zinc physiology. Physiol Rev, 1993. 73(1): p. 79-118.

302. Kristiansen, L.H., et al., Autometallography allows ultrastructural monitoring of zinc in the endocrine pancreas. Histochem Cell Biol, 2001. 115(2): p. 125-9.

303. Rorsman, P., M. Braun, and Q. Zhang, Regulation of calcium in pancreatic alpha- and beta-cells in health and disease. Cell Calcium, 2012. 51(3-4): p. 300-8.

304. Diraison, F., et al., Impact of adenoviral transduction with SREBP1c or AMPK on pancreatic islet gene expression profile: analysis with oligonucleotide microarrays. Diabetes, 2004. 53 Suppl 3: p. S84-91.

305. Wideman, R.D., et al., Improving function and survival of pancreatic islets by endogenous production of glucagon-like peptide 1 (GLP-1). Proc Natl Acad Sci U S A, 2006. 103(36): p. 13468-73.

306. Bellomo, E.A., G. Meur, and G.A. Rutter, Glucose regulates free cytosolic Zn(2)(+) concentration, Slc39 (ZiP), and metallothionein gene expression in primary pancreatic islet beta-cells. J Biol Chem, 2011. 286(29): p. 25778-89.

307. Gee, K.R., et al., Detection and imaging of zinc secretion from pancreatic beta-cells using a new fluorescent zinc indicator. J Am Chem Soc, 2002. 124(5): p. 776-8.

308. Haller, T., et al., The lysosomal compartment as intracellular calcium store in MDCK cells: a possible involvement in InsP3-mediated Ca2+ release. Cell Calcium, 1996. 19(2): p. 157-65.

Page 154: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

154

309. Hutton, J.C., The internal pH and membrane potential of the insulin-secretory granule. Biochem J, 1982. 204(1): p. 171-8.

310. Gerber, P.A., et al., Hypoxia lowers SLC30A8/ZnT8 expression and free cytosolic Zn2+ in pancreatic beta cells. Diabetologia, 2014. 57(8): p. 1635-44.

311. Valentine, R.A., et al., ZnT5 variant B is a bidirectional zinc transporter and mediates zinc uptake in human intestinal Caco-2 cells. J Biol Chem, 2007. 282(19): p. 14389-93.

312. Hou, S., et al., Zn2+ activates large conductance Ca2+-activated K+ channel via an intracellular domain. J Biol Chem, 2010. 285(9): p. 6434-42.

313. Gossen, M. and H. Bujard, Tight control of gene expression in mammalian cells by tetracycline-responsive promoters. Proc Natl Acad Sci U S A, 1992. 89(12): p. 5547-51.

314. Gossen, M., et al., Transcriptional activation by tetracyclines in mammalian cells. Science, 1995. 268(5218): p. 1766-9.

315. Milo-Landesman, D., et al., Correction of hyperglycemia in diabetic mice transplanted with reversibly immortalized pancreatic beta cells controlled by the tet-on regulatory system. Cell Transplant, 2001. 10(7): p. 645-50.

316. Al-Hasani, K., et al., Adult duct-lining cells can reprogram into beta-like cells able to counter repeated cycles of toxin-induced diabetes. Dev Cell, 2013. 26(1): p. 86-100.

317. Mitchell, R.K., et al., Molecular genetic regulation of Slc30a8/ZnT8 reveals a positive association with glucose tolerance. Mol Endocrinol, 2015: p. me20151227.

318. Elayat, A.A., M.M. el-Naggar, and M. Tahir, An immunocytochemical and morphometric study of the rat pancreatic islets. J Anat, 1995. 186 ( Pt 3): p. 629-37.

319. Evans, M.L. and R.S. Sherwin, Brain glucose metabolism and hypoglycaemia. Diabetes Nutr Metab, 2002. 15(5): p. 294-6; discussion 296.

320. Le Marchand, S.J. and D.W. Piston, Glucose decouples intracellular Ca2+ activity from glucagon secretion in mouse pancreatic islet alpha-cells. PLoS One, 2012. 7(10): p. e47084.

321. Rutter, G.A. and T. Tsuboi, Kiss and run exocytosis of dense core secretory vesicles. Neuroreport, 2004. 15(1): p. 79-81.

322. Petersen, A.B., et al., siRNA-mediated knock-down of ZnT3 and ZnT8 affects production and secretion of insulin and apoptosis in INS-1E cells. Apmis, 2011. 119(2): p. 93-102.

323. Smidt, K., et al., SLC30A3 responds to glucose- and zinc variations in beta-cells and is critical for insulin production and in vivo glucose-metabolism during beta-cell stress. PLoS One, 2009. 4(5): p. e5684.

324. Huang, L., et al., Znt7-null mice are more susceptible to diet-induced glucose intolerance and insulin resistance. J Biol Chem, 2012. 287(40): p. 33883-96.

325. Kone, M., et al., LKB1 and AMPK differentially regulate pancreatic beta-cell identity. Faseb j, 2014. 28(11): p. 4972-85.

326. Rutter, G.A. and D.J. Hodson, Beta cell connectivity in pancreatic islets: a type 2 diabetes target? Cell Mol Life Sci, 2015. 72(3): p. 453-67.

327. Cauchi, S., et al., Meta-analysis and functional effects of the SLC30A8 rs13266634 polymorphism on isolated human pancreatic islets. Mol Genet Metab, 2010. 100(1): p. 77-82.

328. Rutter, G.A. and F. Chimienti, SLC30A8 mutations in type 2 diabetes. Diabetologia, 2015. 58(1): p. 31-6.

329. Newsholme, P., et al., Reactive oxygen and nitrogen species generation, antioxidant defenses, and beta-cell function: a critical role for amino acids. J Endocrinol, 2012. 214(1): p. 11-20.

330. Oteiza, P.I., Zinc and the modulation of redox homeostasis. Free Radic Biol Med, 2012. 53(9): p. 1748-59.

Page 155: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

155

331. Fadista, J., et al., Global genomic and transcriptomic analysis of human pancreatic islets reveals novel genes influencing glucose metabolism. Proc Natl Acad Sci U S A, 2014. 111(38): p. 13924-9.

332. da Silva Xavier, G., et al., TCF7L2 regulates late events in insulin secretion from pancreatic islet beta-cells. Diabetes, 2009. 58(4): p. 894-905.

333. Zhou, Y., et al., TCF7L2 is a master regulator of insulin production and processing. Hum Mol Genet, 2014. 23(24): p. 6419-31.

Page 156: Role of the type II diabetes- associated gene SLC30A8 in ...€¦ · 4 Abstract The SLC30A8 locus encodes the Zn2+ transporter ZnT8, whose expression is largely restricted to α-

156

Publications

Gerber, P.A., Bellomo, E.A., Meur, G., Hodson, D.J., Solomou, A., Hollinshead, M., Chimienti, F., Lavallard, V.J., Bosco, D., Hughes, S.J., Johnson, P.R. and Rutter, G.A. Hypoxia alters the expression of the type 2 diabetes-associated Zn2+ transporter Slc30a8/ZnT8 and cytosolic Zn2+ levels in human and rodent islets Diabetologia, 57(8):1635-44 (2014)

Sun, G., da Silva Xavier, G., Gorman, T., Priest, C., Solomou, A., Hodson, D.J., Foretz, M., Viollet, BN., Herrera, P-L., Parker, H., Reimann, F., Gribble, F.M., Migrenne, S., Magnan, C., Marley, A. and Rutter G.A. LKB1 and AMPKa1 are required in pancreatic alpha cells for the normal regulation of glucagon secretion and responses to hypoglycemia. Mol Metab 4(4):277-286 (2015)

Solomou, A., Meur, G., Bellomo, E., Hodson, D.J., Tomas, A., Migrenne Li, S., Philippe, E., Herrera, P., Magnan, C., and Rutter G.A. The zinc transporter Slc30a8/ZnT8 is required in a subpopulation of pancreatic α cells for hypoglycemia-induced glucagon secretion. J Biol Chem 290(35):21432-21442 (2015)

Akalestou, E., Christakis, I, Solomou, A.M., Minnion, J.S., Rutter, G.A. and Bloom, S.R. Proglucagon-derived peptides do not significantly affect acute exocrine pancreas secretion in vitro and in vivo in the rat. Pancreas, in press (2015) Rutter, G.A., Chabosseau, P., Bellomo, E.A., Maret, W., Mitchell, R.K, Hodson, D.J., Solomou, A., and Hu, M. Intracellular zinc in insulin secretion and action: a determinant of diabetes risk? Proc Nutrition Soc. in press (2015)